11.06.2013 Views

Self-Assembly of Synthetic and Biological Polymeric Systems of ...

Self-Assembly of Synthetic and Biological Polymeric Systems of ...

Self-Assembly of Synthetic and Biological Polymeric Systems of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Grupo de Física de Coloides y Polímeros<br />

Departamento de Física de la Materia Condensada<br />

Facultad de Física<br />

Tesis doctoral, 2012<br />

<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> <strong>Synthetic</strong> <strong>and</strong> <strong>Biological</strong><br />

<strong>Polymeric</strong> <strong>Systems</strong> <strong>of</strong> Technological Interest<br />

Josué Elías Juárez On<strong>of</strong>re


Víctor Mosquera Tallón, Catedrático, y Pablo Taboada Antelo, Pr<strong>of</strong>esor Titular, ambos del<br />

Departamento de Física de la Materia Condensada de la Universidad de Santiago de<br />

Compostela,<br />

INFORMAN:<br />

Que el trabajo de investigación titulado “<strong>Self</strong>-assembly <strong>of</strong> <strong>Synthetic</strong> <strong>and</strong> <strong>Biological</strong><br />

<strong>Polymeric</strong> <strong>Systems</strong> <strong>of</strong> Technological Interest”, presentado por Josué Elías Juárez On<strong>of</strong>re ha sido<br />

dirigido por ellos en el Grupo de Física de Coloides y Polímeros del Departamento de Física de<br />

la Materia Condensada de la Universidad de Santiago de Compostela, y reúne los requisitos de<br />

calidad y rigor científicos necesarios para optar al Grado de Doctor en Ciencia y Tecnología de<br />

Materiales.<br />

Para que así conste a los efectos oportunos.<br />

Santiago de Compostela, 6 de febrero de 2012<br />

Fdo. Víctor Mosquera Tallón Fdo. Pablo Taboada Antelo


Agradecimientos<br />

A mi Familia<br />

A mi Padre y Madre


Nuestra vida está dividida en diferentes períodos y en cada uno de ellos siempre se<br />

encuentran personas que influyen de forma esencial en su devenir, regalándonos su amistad y<br />

compartiendo gran parte de su tiempo y conocimiento.<br />

Primero, y de la manera más atenta, agradezco a mis supervisores de este trabajo, Víctor<br />

Mosquera Tallón y Pablo Taboada Antelo, por el conocimiento y soporte científico que me han<br />

dado durante la etapa académica y dirigiendo esta tesis doctoral. También agradecerles el<br />

haber hecho lo “imposible” para que pudiera iniciar (ellos recordarán lo difícil y cómico de la<br />

situación) y desarrollar este Doctorado en Ciencia y Tecnología de Materiales en la Universidad<br />

de Santiago de Compostela.<br />

También me gustaría agradecer al pr<strong>of</strong>esor Stephen Yeates por el apoyo científico y paciencia<br />

durante mi estancia en la Facultad de Química de la Universidad de Manchester. También<br />

quiero agradecer al pr<strong>of</strong>esor Miguel Valdéz de la Universidad de Sonora por su apoyo técnico y<br />

científico.<br />

Por otra parte, también mi más sincero agradecimiento mis compañeros-amigos-colegas del<br />

Grupo de Física de Coloides y Polímeros: Emilio Castro, Sonia Goy, Silvia Barbosa, Adriana<br />

Cambón, Antonio Topete y Manuel Alatorre, por formar parte de este período, compartiendo<br />

su tiempo y conocimiento.<br />

Me siento afortunado en tener a otros muchos compañeros-amigos-colegas y, por eso,<br />

aprovecho para agradecerles su amistad incondicional, y haciendo uso de la poca memoria que<br />

tengo trataré de citarlos en orden cronológico: Felipe Areais, Ceila Fong, Tareixa Regueira, Joel<br />

Pinto, Bea Ordoñez, “Mary Joe”, “Mallegue”, Agueda, “Cío”, Hassan, Irais Batista y José Ríos,<br />

Alej<strong>and</strong>ra Parra, Natalia Hassan (Wua yu Min), Elena Blanco, Paula y Miguel, Liliana, Fern<strong>and</strong>o,<br />

“Toyis” y pido disculpas porque seguro olvidé a alguien. A mis amigos de la residencia Cadarso.<br />

Quiero agradecer también al equipo de microscopía electrónica: Miro, Raquel y Merche, por el<br />

tiempo y paciencia que dedicaron en enseñarme a manejar los equipos de TEM, microscopía<br />

confocal y SEM.<br />

A TODOS, MUCHAS GRACIAS!!!<br />

iii


v<br />

List <strong>of</strong> papers<br />

I. <strong>Self</strong>-assembly process <strong>of</strong> different poly(oxystyrene)-poly(oxyethylene) block copolymers:<br />

spontaneous formation <strong>of</strong> vesicular structures <strong>and</strong> elongated micelles. Juárez, J.; Taboada,<br />

P.; Valdez, M. A.; Mosquera, V. 2008, Langmuir, Vol. 24, pp. 7107-7116.<br />

II. Existence <strong>of</strong> different structural intermediates on the fibrillation pathway <strong>of</strong> human serum<br />

albumin. Juárez, J.; Taboada, P.; Mosquera, V. 2009, Biophysical Journal, Vol. 96, pp.<br />

2353-2370.<br />

III. Influence <strong>of</strong> electrostatic interactions on the fibrillation process <strong>of</strong> human serum albumin.<br />

Juárez, J.; Taboada, P.; Mosquera, V. 2009, J. Phys. Chem. B, Vol. 113, pp. 10521-10529.<br />

IV. Additional supra-self-assembly <strong>of</strong> human serum albumin under amyloid-like-forming<br />

solution conditions. Juárez, J.; Taboada, P.; Goy-López, S.; Cambón, A.; Madec, M. B.;<br />

Yeates, S. G.; Mosquera, V. 2009, J. Phys. Chem. B, Vol. 113, pp. 12391-12399.<br />

V. Surface properties <strong>of</strong> monolayers <strong>of</strong> amphiphilic poly(ethylene oxide)-poly(styrene oxide)<br />

block copolymers. Juárez, J.; Goy-López, S.; Cambón, A.; Valdez, M. A.; Taboada, P.;<br />

Mosquera, V. 2010, J. Phys. Chem. Vol. 114, pp. 15703-15712.<br />

VI. Obtention <strong>of</strong> metallic nanowires by protein biotemplating <strong>and</strong> their catalytic application.<br />

Juárez, J.; Cambón, A.; Goy-López, S.; Topete, A.; Taboada, P.; Mosquera, V. 2010, J.<br />

Phys. Chem. Lett. Vol. 1 pp. 2680-2687.<br />

VII. One-dimensional magnetic nanowires obtained by protein fibril biotemplating. Juárez, J.;<br />

Cambón, A.; Topete, A.; Taboada, P.; Mosquera, V. 2011, Chem. Eur. J. Vol. 17 pp. 7366-<br />

7373.<br />

VIII. Hydration effects on the fibrillation process <strong>of</strong> a globular protein: The case <strong>of</strong> human<br />

serum albumin. Juárez, J.; Alatorre-Meda, M.; Cambón, A.; Topete, A.; Barbosa, S.;<br />

Taboada, P.; Mosquera, V. 2012, S<strong>of</strong>t Matter. DOI:10.1039/C2SM06762E.


vii<br />

Índice general<br />

Agradecimientos ii<br />

Lista de artículos iv<br />

Resumen 1<br />

Abstract 11<br />

CHAPTER 1<br />

<strong>Synthetic</strong> <strong>and</strong> natural polymers<br />

1.1<br />

1.2<br />

1.3<br />

1.4<br />

1.2.1<br />

1.2.2<br />

1.2.3<br />

1.3.1<br />

1.3.2<br />

1.3.3<br />

Chapter 2<br />

Methods<br />

2.1<br />

2.2<br />

2.2.1<br />

2.2.2<br />

2.2.3<br />

Introduction<br />

Classification <strong>of</strong> polymers<br />

<strong>Synthetic</strong> polymers<br />

Biopolymers<br />

Proteins<br />

(Bio)polymer structure<br />

Primary structure<br />

Secondary structure<br />

Tertiary structure<br />

Molecular self-assembly<br />

Introduction<br />

Surface tension<br />

Wilhelmy plate method<br />

Monolayers <strong>and</strong> Langmuir-Blodgett films<br />

Axisymetric drop shape (ADSA)<br />

15<br />

16<br />

17<br />

18<br />

19<br />

19<br />

20<br />

20<br />

22<br />

23<br />

28<br />

28<br />

28<br />

29<br />

31


2.3<br />

2.4<br />

2.5<br />

2.6<br />

2.7<br />

2.8<br />

2.9<br />

2.10<br />

2.11<br />

2.12<br />

2.3.1<br />

2.3.2<br />

2.4.1<br />

2.4.2<br />

2.9.1<br />

2.11.1<br />

2.11.2<br />

2.11.3<br />

2.11.4<br />

2.11.5<br />

Chapter 3<br />

Light scattering<br />

Static light scattering (SLS)<br />

Dynamic light scattering (DLS)<br />

Electron microscopy<br />

Transmission electron microscopy (TEM)<br />

Scanning electron microscopy (SEM)<br />

Atomic force microscopy (AFM)<br />

Fluorescence spectroscopy<br />

Circular dichroism (CD)<br />

Infrared spectroscopy<br />

Rheology<br />

Viscoelasticity<br />

X-ray diffraction<br />

Additional experimental techniques used in fibrillogenesis<br />

studies<br />

Block copolymers<br />

3.1<br />

3.2<br />

3.3<br />

UV-Vis spectroscopy<br />

Polarized light optical microscopy<br />

Emission spectroscopy<br />

Superconducting quantum interference device (SQUID)<br />

Magnetic resonance imaging (MRI)<br />

Laboratory equipment<br />

Introduction<br />

Behaviour <strong>of</strong> block copolymers in solution<br />

Mechanisms <strong>of</strong> morphologic control <strong>and</strong> thermodynamic<br />

stability<br />

32<br />

36<br />

40<br />

44<br />

44<br />

46<br />

47<br />

48<br />

52<br />

54<br />

58<br />

59<br />

64<br />

72<br />

72<br />

75<br />

77<br />

78<br />

80<br />

82<br />

91<br />

93<br />

95<br />

viii


3.4<br />

3.5<br />

Chapter 4<br />

Papers on block copolymers<br />

4.1<br />

4.2<br />

4.3<br />

4.3.1<br />

Chapter 5<br />

Proteins<br />

5.1<br />

5.2<br />

5.3<br />

5.1.1<br />

5.1.2<br />

5.1.3<br />

5.1.4<br />

5.2.1<br />

5.2.2<br />

5.2.3<br />

5.2.4<br />

5.3.1<br />

Block copolymers <strong>of</strong> polyoxyalkylenes in aqueous solution<br />

E12S10, E10S10E10 <strong>and</strong> E137S18E137 block copolymers<br />

<strong>Self</strong>-<strong>Assembly</strong> process <strong>of</strong> different poly(oxystyrene)-<br />

poly(oxyethylene) block copolymers: spontaneous<br />

formation <strong>of</strong> vesicular structures <strong>and</strong> elongated micelles.<br />

Surface properties <strong>of</strong> monolayers <strong>of</strong> amphiphilic<br />

poly(ethylene oxide)-poly(styrene oxide) block copolymers.<br />

Appended papers<br />

Relevant aspects<br />

Protein structure<br />

Primary structure <strong>of</strong> the proteins<br />

Secondary structure <strong>of</strong> the proteins<br />

Tertiary structure <strong>of</strong> the proteins<br />

Quaternary structure <strong>of</strong> the proteins<br />

Protein aggregation<br />

Unfolded state<br />

Intermediate <strong>and</strong> transition state assembles<br />

Native state<br />

Energy l<strong>and</strong>scape: protein aggregation<br />

Amyloid fibrils<br />

Amyloid fibrils hallmarks<br />

96<br />

97<br />

103<br />

113<br />

123<br />

124<br />

127<br />

128<br />

129<br />

131<br />

132<br />

132<br />

133<br />

133<br />

134<br />

135<br />

135<br />

135<br />

ix


5.4<br />

5.5<br />

5.3.2<br />

5.5.1<br />

Chapter 6<br />

Kinetic fibril formation in vitro<br />

Human serum albumin<br />

Papers on amyloid fibrils<br />

6.1<br />

6.2<br />

6.3<br />

6.4<br />

6.5<br />

6.6<br />

6.7<br />

6.8<br />

Amyloid fibrils as biomaterials<br />

Hen egg-white lysozyme<br />

Existence <strong>of</strong> different structural intermediates on the<br />

fibrillation pathway <strong>of</strong> human serum albumin<br />

Influence <strong>of</strong> electrostatic interactions on the fibrillation<br />

process <strong>of</strong> human serum albumin<br />

Additional supra-self-assembly <strong>of</strong> human serum albumin<br />

under amyloid-like-forming solution conditions<br />

Hydration effects on the fibrillation process <strong>of</strong> globular<br />

protein: The case <strong>of</strong> human serum albumin<br />

Obtention <strong>of</strong> metallic nanowires by protein biotemplating<br />

<strong>and</strong> their catalytic application<br />

One-dimensional magnetic nanowires obtained by protein<br />

fibril biotemplating<br />

Relevant aspects on the HSA fibril formation<br />

Relevant aspects on the gold <strong>and</strong> magnetic nanowires<br />

Conclusions 219<br />

References 223<br />

138<br />

139<br />

141<br />

142<br />

147<br />

165<br />

175<br />

185<br />

199<br />

207<br />

215<br />

216<br />

x


Resumen<br />

Desde tiempos ancestrales, las moléculas poliméricas, también conocidas como macromoléculas o<br />

moléculas gigantes, han representado un material útil para el desarrollo de la humanidad.<br />

Actualmente, es difícil imaginar el desarrollo de una sociedad organizada sin el uso de materiales<br />

poliméricos. Por ejemplo, desde hace algunas décadas hasta nuestros días, estos materiales<br />

poliméricos son utilizados para satisfacer diferentes necesidades de gran importancia como son la<br />

construcción de refugios, viviendas (casas, residencias, etc.) y vestimenta; dependiendo del uso<br />

que se requiera, se emplearán materiales poliméricos sintéticos (tales como el poliéster,<br />

polietileno, polipropileno, teflón, resinas, entre otros) y/o naturales, también conocidos como<br />

biopolímeros (madera, pieles, gomas y resinas naturales, algodón seda, lana). Por otra parte,<br />

diferentes macromoléculas desempeñan una función de gran importancia biológica.<br />

Biomacromoléculas como los ácidos ribonucleicos (ADN y ARN), proteínas y polisacáridos son de<br />

vital importancia en distintos procesos bioquímicos para el crecimiento y desarrollo de un<br />

organismo vivo; además, proteínas y polisacáridos son utilizados por los organismos vivos como<br />

soportes estructurales de células, tejidos y órganos.<br />

El uso de los materiales poliméricos no sólo está limitado a los usos descritos anteriormente. Hoy<br />

en día una gran variedad de materiales poliméricos, tanto sintéticos como naturales, están siendo<br />

utilizados en diferentes áreas tecnológicas como la ciencia de materiales, medicina clínica o<br />

industria farmacéutica, entre otras. Además, los procesos de síntesis química, hacen posible<br />

diseñar una macromolécula con propiedades fisicoquímicas deseadas. Con este fin, diversos<br />

grupos de investigación han sintetizado diferentes tipos de copolímeros con carácter anfifílico. El<br />

diseño de la arquitectura molecular del copolímero puede ser tan sencillo como un polímero<br />

lineal, o tan complejo como un copolímero de injerto o tipo estrella; la arquitectura molecular<br />

puede controlarse por medio de procedimientos de síntesis química.<br />

Nuestro trabajo de investigación se divide en dos etapas: la primera parte se enfocó a la<br />

caracterización estructural y de las propiedades físico-químicas de tres copolímeros de bloque con<br />

potencial interés para ser empleados como sistemas de liberación farmacológica; mientras que en<br />

la segunda parte se ha realizado un estudio del proceso de autoensamblaje en la forma de fibras<br />

1


amiloides de la proteína albúmina de suero humano (HSA), su caracterización estructural y de sus<br />

propiedades físico-químicas, y su uso potencial como biomateriales.<br />

Debido a que parte de este trabajo consiste en la caracterización de tres copolímeros de bloque,<br />

describiremos brevemente este tipo de macromoléculas. Para un copolímero, el “esqueleto” de la<br />

cadena macromolecular está constituido por más de un monómero o unidad de repetición. Por lo<br />

tanto, la estructura molecular de un copolímero de bloque consiste en la unión química de dos o<br />

más macromoléculas de homopolímeros diferentes. De acuerdo al número de bloques (unidades<br />

homopoliméricas) estos polímeros se denominan copolímeros dibloque, tribloque o multibloque.<br />

Generalmente, estos homopolímeros son de diferente naturaleza (hidrófilo o hidrófobo)<br />

proporcion<strong>and</strong>o a la macromolécula un carácter anfifílico.<br />

Bajo ciertas condiciones, los copolímeros pueden adoptar una configuración tipo gusano o<br />

espagueti, mientras que para otras, el polímero puede auto-asociarse hasta formar<br />

nanoestructuras regulares. Por ejemplo, si un copolímero de bloque se encuentra en disolución<br />

acuosa, en donde el agua es un solvente selectivo para uno de los bloques, los segmentos<br />

hidrófobos de las macromoléculas tienden a auto-asociarse, form<strong>and</strong>o estructuras específicas,<br />

evit<strong>and</strong>o el contacto con las moléculas del disolvente. La auto-asociación origina un amplio rango<br />

de comportamientos de fases, en los que se incluye la formación de micelas, nanovarillas,<br />

vesículas o gr<strong>and</strong>es agregados. Estas estructuras poliméricas presentan, en particular, un especial<br />

interés en el área biomédica. Así, las estructuras micelares y vesiculares pueden ser utilizados<br />

como sistemas de liberación de fármacos y proteínas, así como en el desarrollo de nanoreactores y<br />

de biosensores.<br />

Los copolímeros de bloque de oxialquilenos están constituidos, generalmente, por el bloque<br />

hidrófilo polióxido de etileno (OCH2CH2 es la unidad de óxido de etileno, EO) y por un segundo<br />

bloque hidrófobo que puede ser: un bloque de polióxido de propileno (PO), un bloque de<br />

polióxido de estireno (SO) u otro tipo de bloque hidrófobo. En medio acuoso, la estructura de los<br />

agregados que forman estos copolímeros de bloque está gobernada por el balance de las fuerzas<br />

de interacción entre los bloques moleculares que constituyen el esqueleto del copolímero con las<br />

moléculas de agua, y por las interacciones intermoleculares entre las cadenas del copolímero (ya<br />

2


que las cadenas hidrófilas se exp<strong>and</strong>en evit<strong>and</strong>o las interacciones entre segmentos que no sean<br />

favorables, mientras que las hidrófobas se contraen para minimizar el contacto con las moléculas<br />

del disolvente). Este tipo de interacciones pueden modificarse por medio de factores físicos y, de<br />

esta manera, se puede controlar la arquitectura de los agregados. Los factores físicos que se varían<br />

con más frecuencia son: el peso molecular, la longitud relativa de los bloques, la estructura de los<br />

bloques hidrófobos, las propiedades en disolución, etc. De esta forma, se ha observado que en<br />

disoluciones acuosas diluidas estos copolímeros de bloque anfifílicos se asocian para formar<br />

micelas esféricas. Al aumentar la concentración del copolímero se obtienen mes<strong>of</strong>ases liotrópicas<br />

cristalinas (geles), en donde las micelas se encuentran altamente empaquetadas en una estructura<br />

cúbica centrada en las caras (fcc) o centrada en el cuerpo (bcc).<br />

En el presente trabajo, se estudió el proceso de micelización, gelificación y la estructura de los<br />

agregados de tres copolímeros de poli(óxido de etileno)-poli(óxido de estireno): EO12SO10,<br />

EO10SO10EO10, y EO137SO18EO137, donde los subíndices indican la longitud de cada bloque. De estos<br />

polímeros, dos tienen la misma longitud del bloque de poli(óxido de estireno) (EO12SO10 y<br />

EO10SO10EO10,), mientras que la longitud del bloque EO para el tercer copolímero es mucho más<br />

largo (EO137SO18EO137). Es importante mencionar que estos copolímeros de bloque fueron<br />

sintetizados en colaboración con los Pr<strong>of</strong>s. David Attwood y Colin Booth de la en la School <strong>of</strong><br />

Pharmacy <strong>and</strong> Pharmaceutical Sciences y en la School <strong>of</strong> Chemistry <strong>of</strong> the University <strong>of</strong><br />

Manchester, respectivamente. La caracterización de estos copolímeros de bloque se realizó en dos<br />

fases: La primera parte consistió en el análisis de las propiedades estructurales y físico-químicas de<br />

los copolímeros en el seno de disoluciones acuosas, mediante diversas técnicas experimentales<br />

tales como: medidas de tensión superficial, medidas de dispersión de luz láser (DLS y SLS),<br />

microscopía óptica de luz polarizada (POM), microscopía de transmisión electrónica (TEM) y<br />

microscopía electrónica de barrido (SEM). Por medio de las medidas experimentales para los<br />

copolímeros EO12SO10, EO10SO10EO10, y EO137SO18EO137, combinado con modelos teóricos (según los<br />

datos experimentales obtenidos) ya publicados, se determinaron la concentración micelar crítica,<br />

así como propiedades micelares (tamaño de las micelas y número de agregación) y vesiculares<br />

(para el caso del copolímero EO12SO10); imágenes de los agregados estructurados y el<br />

comportamiento reológico. En este trabajo, en particular se observó por primera vez la posibilidad<br />

de que los copolímeros de poli(óxido de etileno)-poli(óxido de estireno) pueden formar<br />

estructuras vesiculares, sin necesidad de aportes externos de energía. Estas estructuras<br />

3


vesiculares son el resultado de la auto-asociación del copolímero EO12SO10 en disolución acuosa, y<br />

se encuentran en coexistencia con estructuras micelares esféricas. Este resultado se confirmó por<br />

dispersión de luz láser, TEM y cryo- SEM. El tamaño de las vesículas varía entre 60 y 500 nm. Para<br />

los copolímeros EO10SO10EO10 y EO137SO18EO137 se observaron estructuras en la forma de micelas<br />

alargadas y esféricas, respectivamente. Compar<strong>and</strong>o el alto número de agregación obtenido para<br />

las micelas del copolímero EO10SO10EO10 por dispersión de luz láser con el número máximo de<br />

agregación teórico que correspondería a una micela esférica, se puede concluir que la geometría<br />

micelar no se corresponde con una geometría esférica, pero sí con una geometría más alargada.<br />

Este resultado es confirmado por medio de imágenes obtenidas por TEM. La prueba de inversión<br />

de tubo se utilizó para definir las fronteras que dividen la fase de un gel bl<strong>and</strong>o con la fase de un<br />

gel duro de las disoluciones del copolímero EO137SO18EO137. De acuerdo con los resultados<br />

obtenidos por reología, los copolímeros EO12SO10 y EO10SO10EO10 no forman geles en el rango de<br />

concentraciones estudiados en este trabajo.<br />

Por otra parte, la capacidad de adsorción y de auto-organización molecular en una interface de un<br />

copolímero de bloque para formar nanoestructuras bien definidas en dos dimensiones permiten<br />

que estos materiales poliméricos puedan utilizarse como estabilizantes coloidales, o como agentes<br />

recubridores y estabilizantes de espumas y emulsiones.<br />

Así, un amplio rango de estudios han analizado el comportamiento interfacial de copolímeros de<br />

poli(óxido de etileno)-poli(óxido de propileno)-poli(óxido de etileno) (EO-PO-EO) ó “Pluronics”,<br />

copolímeros de estrella de EO-PO-EO, también conocidos como “Tetronics” y de los copolímeros<br />

poli(óxido de butileno)-poli(óxido de etileno) (BO-EO). Sin embargo, pocos estudios existen sobre<br />

la caracterización en la interfase de los copolímeros de poli(óxido de etileno)-poli(óxido de<br />

estireno). El análisis de sus propiedades de adsorción es interesante ya que puede proporcionar<br />

importante información sobre su estabilidad intefacial y posterior agregación. Así, en el presente<br />

trabajo, se estudiaron las propiedades de adsorción en la interface aire-agua y en la interface<br />

clor<strong>of</strong>ormo-agua de los copolímeros EO12SO10, EO10SO10EO10 y EO137SO18EO137. El trabajo de<br />

investigación se desarrolló mediante las técnicas experimentales: tensiómetro de gota ,<br />

monocapas o películas de Langmuir-Blodgett y microscopía de fuerza atómica (AFM). Las cinéticas<br />

de adsorción de las disoluciones copoliméricas, en ambas interfaces, se determinaron por la<br />

técnica de gota suspendida. Los resultados de tensión superficial muestran que la adsorción en la<br />

4


interface aire-agua de los copolímeros estudiados disminuye conforme se incrementa la<br />

hidr<strong>of</strong>obicidad del copolímero. Los datos obtenidos mediante reología interfacial de la película<br />

copolimérica adsorbida en la interface aire-agua muestran que la capa adsorbida en la interface se<br />

comporta como un sólido en todo el rango de frecuencias analizadas, mientras que la capa<br />

polimérica adsorbida en la interface clor<strong>of</strong>ormo-agua se comporta como un fluido viscoso para el<br />

mismo rango de frecuencias. Las isotermas de compresión, presión superficial vs área ( vs A), y<br />

las monocapas de Langmuir-Blodgett se obtuvieron con una balanza de Langmuir-Blodgett. Las<br />

diferencias en las isotermas de Langmuir obtenidas para los tres copolímeros de bloque tienen su<br />

origen en la diferente relación entre los bloques hidrófobo/hidrófilo. Bajo esta consideración, el<br />

copolímero EO137SO18EO137 muestra una isoterma de adsorción con cuatro estados característicos,<br />

mientras que en las isotermas de adsorción de los otros dos copolímeros (EO12SO10 y<br />

EO10SO10EO10), sólo se observan dos regiones. Las monocapas de Langmuir-Blodgett se obtuvieron<br />

a dos diferentes valores de presión superfcial (5 y 11 mN/m), y sus imágenes topográficas fueron<br />

analizadas por AFM. Para ambas presiones, las imágenes topográficas revelan la existencia de<br />

micelas circulares en la superficie de la película. Sin embargo, el tamaño de las micelas disminuye<br />

con el aumento de la presión interfacial y, a su vez se observa un incremento del espesor de la<br />

monocapa.<br />

En la segunda etapa de nuestro trabajo, se estudiaron los procesos de ensamblaje molecular en la<br />

forma de fibras en disolución acuosa de la proteína seroalbúmina humana (HSA, sus siglas en<br />

inglés), bajo diferentes condiciones de temperatura, pH y fuerza iónica; y la posible aplicación de<br />

estas nanoestructuras en diversas aplicaciones como catalizadores o elementos de contraste de<br />

imagen.<br />

In vivo, la agregación incontrolada de ciertas proteínas está relacionada directamente con<br />

diferentes dolencias que afectan la salud humana. Uno de los productos de mayor interés<br />

originados por la inadecuada agregación proteica son las fibras amiloides. Estas estructuras<br />

fibrilares se han detectado en con más de 20 enfermedades (neurodegenerativas o no) tales como<br />

el Alzheimer, Huntington, el mal de Parkinson, o la diabetes tipo II, entre otras. La capacidad de<br />

fibrilación es independiente de la estructura nativa de la proteína, no obstante, la estructura<br />

proteica primaria aparece como un factor importante en las primeras fases del proceso de<br />

formación de las fibras amiloides, en su cinética de formación y en su estabilidad. La fibrilación se<br />

5


origina debido a condiciones ambientales en las que se promueve la desnaturalización parcial o<br />

completa de la estructura nativa de la proteína, por lo que para una proteína o cadena<br />

polipeptídica, la formación de la fibra amiloide representa una estructura estable alternativa a la<br />

estructura nativa.<br />

Las fibras tipo amiloide presentan características termodinámicas y estructurales bien definidas<br />

Estas fibras se caracterizan, en general, por ser lineales, rígidas y sin ramificaciones, con una<br />

anchura que varía entre los 5-12 nm, tal como se ha observado por microscopía de transmisión<br />

electrónica; además, tienen la propiedad de enlazarse a marcadores orgánicos específicos, como<br />

Congo Red y Ti<strong>of</strong>lavina T. Muestran también un patrón común de difracción de rayos-X<br />

característico para una estructura -cruzada. Como se mencionó anteriormente, las fibras de<br />

amiloide son responsables de distintas enfermedades como el mal de Parkinson, deficiencia de 1-<br />

tripsina, Alzheimer y encefalopatías espongiformes. El origen de estas enfermedades está asociado<br />

a la conversión de proteínas monoméricas a agregados proteicos insolubles, los cuales se<br />

depositan en los tejidos/órganos del cuerpo.<br />

La proteína plasmática más abundante del sistema circulatorio es la albúmina (representa<br />

aproximadamente del 52 al 60 % de las proteínas plasmáticas), cuyas funciones son la regulación<br />

de la presión osmótica sanguínea y en el transporte y almacenamiento de moléculas endógenas y<br />

exógenas (hormonas, iones, ácidos grasos, drogas). A nivel industrial, la albúmina tiene un rol<br />

importante en la tecnología médica, de alimentos, fármacos y en el desarrollo de biosensores<br />

debido a sus fuertes propiedades de adsorción. La HSA es una proteína globular formada por una<br />

sola cadena polipeptídica. La estructura de esta proteína está constituida por 585 residuos de<br />

aminoácidos, con un peso molecular de aproximadamente 66,500 g/mol. La estructura secundaria<br />

de la HSA consta principalmente de hélices (60 %); además presenta 17 puentes disulfuro, los<br />

que le confiere una gran estabilidad estructural. La estructura globular de la HSA está compuesta<br />

por tres dominios principales y seis subdominios unidos por puentes disulfuro. Debido a su<br />

importancia fisiológica como proteína de transporte y reguladora de la presión osmótica, y a su<br />

facilidad de agregación in vitro, la HSA se puede considerar como un modelo adecuado para<br />

realizar estudios sobre los mecanismos de agregación proteica, con el fin de obtener una<br />

aproximación y un mejor entendimiento de los mecanismos de agregación proteica asociada a la<br />

amiloidogénesis. Ya que la estructura nativa de la HSA carece de cualquier predisposición a formar<br />

6


fibras amiloides, la agregación de la albúmina sérica se puede promover cambi<strong>and</strong>o las<br />

condiciones de disolución (tales como de pH, altas temperaturas y por medio de agentes químicos<br />

desnaturalizantes), favoreciendo la desestabilización de la estructura nativa de la HSA.<br />

Así, en este trabajo se analizó la predisposición a la fibrilación de la HSA bajo diferentes<br />

condiciones de disolución (pH, temperatura, fuerza iónica, composición del disolvente), así como<br />

las propiedades físicas y estructurales de las fibras amiloidales resultantes. Las cinéticas de<br />

agregación, los cambios conformacionales durante la auto-asociación de los monómeros de<br />

albúmina y las estructuras de las diferentes estructuras intermedias en la ruta de fibrilación se<br />

determinaron por fluorescencia de Ti<strong>of</strong>lavina T (ThT) y absorbencia del Congo red; dicroísmo<br />

circular (CD); fluorescencia intrínseca de la HSA; espectroscopía de infra-rojo (FTIR); difracción de<br />

rayos-X; TEM; SEM y AFM. El proceso de la fibrilación de la HSA se prolonga durante varios días de<br />

incubación, y se lleva a cabo sin la presencia de una fase de demora; excepto cu<strong>and</strong>o las<br />

disoluciones de albúmina se incubaron a pH ácido, temperatura ambiente y en ausencia de<br />

electrólito. La ausencia de la fase de demora se debe al hecho de que la agregación inicial está<br />

dirigida por el mecanismo denominado “cuesta abajo”, que no requiere de una alta organización<br />

estructural ni de la formación de núcleos proteicos inestables. El proceso de fibrilación de la HSA<br />

está acompañada por un incremento progresivo de la estructura secundaria hoja-β y estructuras<br />

ordenadas al azar; esta transición se lleva a cabo a expensas de la estructura α-hélice, como se<br />

observó mediante espectroscopía de fluorescencia de ThT, de CD y de FTIR. De acuerdo a las<br />

distintas condiciones de incubación, la estructura nativa de la proteína se altera de distinta<br />

manera/grado, implic<strong>and</strong>o una variación en el contenido de las diferentes estructuras secundarias;<br />

estas variaciones implican la aparición de diversos agregados estructurales que se forman en el<br />

proceso de agregación fibrilar. Estos agregados pueden tomar la forma de: clústeres oligoméricos<br />

globulares, estructuras rígidas esféricas y agregados en forma de anillos. La formación fibrilar<br />

puede tener lugar a través de la asociación de los intermedios oligoméricos mencionados,<br />

conduciendo a la formación de prot<strong>of</strong>ibras y, posteriormente, fibras. Estas fibras resultantes son<br />

alargadas, con ondulaciones. La longitud de las fibras obtenidas depende de las condiciones de<br />

incubación. Probablemente las hebras-β que componen la estructura de hoja-β en las fibras de<br />

HSA se encuentran en un arreglo antiparalelo. La aparición de fibras amiloides maduras con<br />

estructuras morfológicas más complejas tiene lugar a largos periodos de incubación (fibras largas y<br />

rígidas, estructuras planas y alargadas, etc.). La tendencia de los oligómeros proteicos a alinearse<br />

7


con las fibras vecinas más cercanas puede resultar en el crecimiento y maduración de las fibras<br />

rígidas. Además, estas fibras pueden asociarse en otras estructuras supramoleculares tales como<br />

esferulitas e incluso geles fibrilares. La presencia de esferulitas de fibras amiloides se confirmó por<br />

POM, microscopía confocal y TEM. La HSA tiene la capacidad de formar geles fibrilares<br />

entrecruzados, los cuales se forman por medio de interacciones intermoleculares no específicas a<br />

valores de pH que se encuentran lejos del punto isoeléctrico de la HSA, cu<strong>and</strong>o se sobrepasa un<br />

umbral de concentración de proteínas y/o de fuerza iónica (punto de gelificación), tal como se<br />

observó por reología. Por otra parte, cu<strong>and</strong>o la HSA se encuentra cerca de su punto isoeléctrico, el<br />

proceso de agregación proteica es tan rápido, que no permite una organización estructural<br />

substancial para permitir la formación de estructuras ordenadas, form<strong>and</strong>o así un gel de tipo<br />

particulado.<br />

A pesar de la importancia de las fibras amiloides en el área clínica, actualmente existen otras áreas<br />

científicas que prestan especial interés en el estudio de las propiedades estructurales y de<br />

agregación fibrilar, en la búsqueda de nuevos materiales y posibles aplicaciones en campos de<br />

ingeniería y nanotecnología. El desarrollo de estructuras a escala mesoscópica es de gran<br />

importancia en el desarrollo de dispositivos opto-electrónicos, como biosensores, materiales<br />

biocompatibles, etc. Usualmente, los materiales nanoestructurados se obtienen a través de dos<br />

estrategias: La primera, llamada “top-down”, consiste en la reducción de tamaño de un material<br />

más gr<strong>and</strong>e; sin embargo, mediante esta estrategia se obtienen materiales en escala micrométrica<br />

debido a que la reducción del tamaño sólo se puede realizar hasta donde las propiedades del<br />

tamaño del material lo permita. En la segunda estrategia, “bottom-up”, se utilizan moléculas<br />

pequeñas, como las proteínas, péptidos, fosfolípidos y ácidos nucleicos (ADN y RNA) como<br />

elementos constituyentes, que bajo condiciones controladas, permiten obtener estructuras bien<br />

ordenadas gracias a las propiedades de auto-ensamblaje que poseen este tipo de biomoléculas.<br />

Las propiedades de auto-asociación y ensamblaje que tienen estos biomateriales permiten, bajo<br />

condiciones controladas, obtener complejos nanoestructurados definidos. Las proteínas son una<br />

alternativa muy atractiva para la construcción de estas nanoestructuras, ya que además de tener<br />

un tamaño físico adecuado una gran estabilidad (comparado con los estructutras basadasen ácidos<br />

nucleicos y fosfolípidos), y unas excelentes propiedades de adsorción a sustratos como vidrio,<br />

óxido de silicio, u oro; por otra parte las proteínas presentan una serie de funciones que pueden<br />

8


ser acopladas, por ejemplo, a circuitos electrónicos durante la construcción de nanodispositivos<br />

gracias a su alta flexibilidad, aún cu<strong>and</strong>o estos materiales son no conductores, pueden recubrirse<br />

con metales conductores.<br />

En la parte final de nuestro trabajo hemos analizado el potencial de las fibras amiloides como<br />

nuevos nanomateriales con potenciales aplicaciones en distintos campos como la catálisis o la<br />

biomedicina Bajo este contexto, en perime lugar se obtuvieron nanocables de oro por medio de<br />

un proceso de crecimiento de “semillas” de oro utiliz<strong>and</strong>o fibras de la proteína lisozima como<br />

bioplantillas. Se decidió emplear lisozima pues sus fiobras han sido muy estudiadas y7 presnetan<br />

una geomtería recta sin ramificaciones, muy adecuada para el objetivo de la construcción de un<br />

nanocable metálico. El grado de metalización de la fibra se controló por la adicción secuencial de<br />

sal de oro hasta alcanzar un recubrimiento completo de la fibra. Los híbridos orgánico-metálicos<br />

obtenidos se utilizaron como catalizadores en la reducción del p-nitr<strong>of</strong>enol a p-amin<strong>of</strong>enol.<br />

Por otra parte, se obtuvieron nanocables magnéticos utiliz<strong>and</strong>o como bioplantillas las fibras de<br />

lisozima y de HSA. La obtención de estos materiales híbridos se realizó por medio de una reacción<br />

de precipitación in situ de la mezcla de sales de hierro Fe +2 :Fe +3 en una relación 1:2, en presencia<br />

de NH4OH, siendo el producto de esta reacción nanopartículas de magnetita. Las propiedades<br />

magnéticas de los híbridos se determinaron mediante SQUID (por sus siglas en inglés:<br />

superconducting quantum interference device), y se compararon con las propiedades magnéticas<br />

de nanopartículas libres de magnetita, observándose una mejora sustancial en sus propiedades<br />

magnéticas. Finalmente, la posible utilidad de estos materiales biohíbridos como agentes de<br />

contraste se corroboró mediante imágenes de resonancia magnética nuclear a diferentes<br />

proporciones de dopado de la fibra proteica.<br />

9


Abstract<br />

In the present work, we investigated the micellization, gelation <strong>and</strong> the structure <strong>of</strong> aggregates <strong>of</strong><br />

three poly(ethylene oxide)-poly(styrene oxide)-poly(ethylene oxide) block copolymers (EO12SO10,<br />

EO10SO10E10 <strong>and</strong> EO137SO18EO137, where E represents the oxyethylene unit, S the oxystyrene unit,<br />

<strong>and</strong> the subscripts correspond to the number <strong>of</strong> monomeric unit constituting the polymeric chain)<br />

in solution. We also investigated the adsorption process <strong>and</strong> the surface properties <strong>of</strong> these block<br />

copolymers at the air-water (a/w) <strong>and</strong> chlor<strong>of</strong>orm-water (c/w) interfaces. Since these copolymers<br />

possess a more hydrophobic middle block –the styrene oxide unit, they should gives rise to more<br />

efficient drug delivery systems, in which geometry might play a key role for drug solubilisation <strong>and</strong><br />

cell uptake. For these reasons, a detailed characterization <strong>of</strong> the physicochemical properties <strong>of</strong><br />

these copolymers is required.<br />

In this work, the spontaneous formation <strong>of</strong> vesicles by the block copolymer EO12SO10 without the<br />

need <strong>of</strong> external energy input has been observed In aqueous bulk solution. This is the first time<br />

that vesicle formation for styrene oxide based-block copolymers is reported. These vesicular<br />

structures are in coexistence with spherical micelles, as confirmed by dynamic light scattering<br />

(DLS), polarized optical microscopy <strong>and</strong> transmission <strong>and</strong> cryo-scanning electron microscopies.<br />

Meanwhile, for block copolymers EO10SO10E10 <strong>and</strong> EO137SO18EO137 only one species is found in<br />

solution, assigned to elongated <strong>and</strong> spherical micelles, respectively. The adsorption kinetics <strong>of</strong><br />

these block copolymers at both (a/w) <strong>and</strong> (c/w) interfaces were determined by the pendant drop<br />

method. Measurements <strong>of</strong> their interfacial rheological behaviour showed that the adsorption<br />

layer at the (a/w) interface manifests solid-like properties in the whole accessible frequency range,<br />

whereas a viscous fluid-like behaviour is displayed at the (c/w) interface. The adsorption isotherm<br />

for the block copolymer EO137SO18EO137 displays the four classical regions (pancake, mushroom,<br />

brush, <strong>and</strong> condensed states), whereas for EO12SO10 <strong>and</strong> EO10SO10E10 copolymers only two regions<br />

were observed. Probably, these observed differences in the monolayer isotherms arose from their<br />

different hydrophobic/hydrophilic block ratios <strong>and</strong> block lengths.<br />

On the other h<strong>and</strong>, the fibrillation propensity <strong>of</strong> the protein human serum albumin (HSA) was<br />

analysed under different solution conditions (varying pH, temperature, ionic strength <strong>and</strong> solvent<br />

11


composition). Fibrillation is an important self-assembly mechanism <strong>of</strong> proteins which allows them<br />

to achieve an energy state even lower than their own native state. Also, the presence <strong>of</strong> protein<br />

fibrils has been related to the appearance <strong>and</strong> development <strong>of</strong> more than 20 neurodegenerative<br />

<strong>and</strong> non-neurodegenerative diseases, known as amyloid diseases. Hence, a pr<strong>of</strong>ound knowledge<br />

about the fibrillation mechanisms <strong>and</strong> the structure <strong>of</strong> the resulting protein assemblies is required<br />

to develop new therapeutical strategies, but also to exploit the special structural properties <strong>of</strong><br />

such protein aggregates in new technological fields.<br />

In this way, the kinetics <strong>of</strong> HSA fibril formation under different solution conditions <strong>and</strong> the<br />

structures <strong>of</strong> resulting fibrillar aggregates were determined. It was observed that fibril formation is<br />

largely affected by electrostatic shielding: At physiological pH, fibrillation is more efficient <strong>and</strong><br />

faster in the presence <strong>of</strong> up 50 mM NaCl. In contrast, under acidic conditions a continuous<br />

progressive enhancement <strong>of</strong> HSA fibrillation is observed as the electrolyte concentration in<br />

solution increases. The fibrillation process is accompanied by a progressive increase in β-sheet <strong>and</strong><br />

unordered conformations at the expense <strong>of</strong> α-helix one as revealed by CD, FT-IR <strong>and</strong> tryptophan<br />

fluorescence spectra. The presence <strong>of</strong> structural intermediates in the aggregation pathway, such<br />

as oligomeric clusters (globules), bead-like structures, <strong>and</strong> ring-shaped aggregates suggest that<br />

fibril formation takes place by the competent-association <strong>of</strong> these aggregates. The resultant fibrils<br />

are elongated but curly, <strong>and</strong> differ in length depending on solution conditions. In addition, very<br />

long incubation times lead to a complex morphological variability <strong>of</strong> amyloid fibrils, such as long<br />

straight fibrils or flat-ribbon structures, amongst other. In addition, the formations <strong>of</strong> ordered<br />

aggregates <strong>of</strong> amyloid fibrils, such as spherulites, which possess a radial arrangement <strong>of</strong> the fibrils<br />

around a disorganized protein core, or fibrillar gels were also observed. All these observations lead<br />

to the conclusion that HSA can be used as a suitable model to study protein fibrillogenesis.<br />

On the other h<strong>and</strong>, the utility <strong>of</strong> the obtained HSA fibrils as building blocks to construct new <strong>and</strong><br />

useful nanomaterials as tested. We reported a method to obtain metallic Au nanowires by using<br />

lysozyme protein fibrils as bioscaffolds to generate a complete gold coating layer on the<br />

biotemplate surface by the attachment <strong>of</strong> gold seeds <strong>and</strong> subsequent overgrowth <strong>of</strong> the coating<br />

layer by sequential addition <strong>of</strong> gold salt growth solution (seeded-mediated mechanism). We<br />

decided to use lysozyme fibrils since these display a straight <strong>and</strong> uniform core, which is more<br />

suitable for the generation <strong>of</strong> nanowires. To obtain full coverage <strong>of</strong> the protein fibril, suitable<br />

12


[Au]/[protein] molar ratio <strong>and</strong> number <strong>of</strong> sequential additions <strong>of</strong> Au ions must be chosen. The<br />

hybrid metallic fibrils have been proved to be useful as catalytic substrates, provided their great<br />

catalytically activity when incorporated in the reduction reaction <strong>of</strong> p-nitrophenol to p-<br />

aminophenol by NaBH4. On the other h<strong>and</strong>, we also reported a method to obtain magnetic Fe3O4<br />

nanowires by using HSA <strong>and</strong> lysozyme protein fibrils as bioscaffolds to generate a magnetic Fe3O4<br />

coating layer on the biotemplate surface by the in situ controlled Fe nanoprecipitation method.<br />

The magnetic properties <strong>of</strong> the magnetic nanowires have been tested by using a superconducting<br />

quantum interference device (SQUID), <strong>and</strong> their potential applicability as MRI contrast agents was<br />

analysed by means <strong>of</strong> magnetic resonance imaging at different Fe doping concentration in<br />

phantoms.<br />

13


Contents<br />

1.1<br />

1.2<br />

1.3<br />

1.4<br />

CHAPTER 1<br />

SYNTHETIC AND NATURAL POLYMERS<br />

1.2.1<br />

1.2.2<br />

1.2.3<br />

1.3.1<br />

1.3.2<br />

1.3.3<br />

1.1.- Introduction<br />

Introduction<br />

Classification <strong>of</strong> polymers<br />

<strong>Synthetic</strong> polymers<br />

Biopolymers<br />

Proteins<br />

(Bio)polymer structure<br />

Primary structure<br />

Secondary structure<br />

Tertiary structure<br />

Molecular self-assembly<br />

Since many years ago, polymeric molecules have progressively gained a key role in human´s<br />

life <strong>and</strong> activity. Many examples about the use <strong>of</strong> polymers can be easily found in our daily<br />

activity in different fields such as clothing, health care, packing or waste water treatments;<br />

from the early times, humans have taken advantage <strong>of</strong> polymeric materials in the form <strong>of</strong><br />

wood, leather, natural gums <strong>and</strong> resins, cotton fibres, wool <strong>and</strong> silk (1).<br />

<strong>Polymeric</strong> molecules found in Nature also possess very important biological functions as in the<br />

case <strong>of</strong> nucleic acids, proteins or carbohydrates. These biomacromolecules take part in<br />

different biochemical processes <strong>and</strong> are necessary in anabolic <strong>and</strong> catabolic processes;<br />

furthermore, they fulfill an essential structural role during the development <strong>and</strong> growth <strong>of</strong><br />

living organisms as evidenced, for example, in plants, which are mainly built from cellulose<br />

fibres, or animals, composed <strong>of</strong> linear protein materials such collagen in skin, sinew <strong>and</strong> bone,<br />

myosin in muscles, <strong>and</strong> keratin in nails <strong>and</strong> hair. Polymers are giant molecules built up by<br />

repetitive covalent unions <strong>of</strong> small chemical units known as monomers. Polymers usually have<br />

high molecular weights <strong>and</strong>, for this reason, are also called “macromolecules”. They can be<br />

classified in two main groups regarding their origin: synthetic <strong>and</strong> biological ones. In general,<br />

15<br />

16<br />

17<br />

18<br />

19<br />

19<br />

20<br />

20<br />

22<br />

23<br />

15


synthetic polymers are obtained from processes developed in laboratories while biological<br />

polymers (biopolymers) are found in nature, but many <strong>of</strong> them can be also synthetically<br />

synthesized (1)-(3).<br />

From a historical perspective, first attempts to obtain lab-synthesized polymers dated from<br />

1868 with the synthesis <strong>of</strong> cellulose nitrate by the chemical modification <strong>of</strong> natural polymers<br />

(pyroxin made from cotton <strong>and</strong> nitric acid with camphor). However, in 1909, Leo Hendrick<br />

Baekel<strong>and</strong> synthesized a plastic material, the bakelite, which can be considered as the true<br />

first synthetic polymer. From 1920s, different polymers <strong>of</strong> commercial importance were<br />

synthesised; however, the main interest was focused on controlling their chemical<br />

composition along the synthetic process, <strong>and</strong> no cautions about a suitable control <strong>of</strong> the final<br />

molecular structure were considered since the synthesis <strong>of</strong> macromolecules was an empirical<br />

work. In the middle <strong>of</strong> 20 th century, chemists <strong>and</strong> physicists provided new methods to establish<br />

the structural description <strong>of</strong> polymeric molecules. In this regard, Herman Staudinger´s, a<br />

German chemist who introduced the concept <strong>of</strong> macromolecule, <strong>and</strong> W. T. Carother´s, an<br />

American chemist who classified the polymers according to their preparation method (i.e.,<br />

addition <strong>and</strong> condensation polymers), works deserve special attention.<br />

From a general point <strong>of</strong> view, polymer molecules show considerable flexibility due to rotation<br />

around molecular bonds. In solution, the shape <strong>of</strong> these macromolecules alters continuously<br />

due to thermal energy so they can be treated as r<strong>and</strong>om coils, even when the rotation around<br />

molecular bonds does allow complete flexibility, while steric <strong>and</strong> excluded volume effects<br />

oppose the formation <strong>of</strong> this kind <strong>of</strong> configuration. For these reasons, dissolved linear polymer<br />

molecules will tend to be more extended than r<strong>and</strong>om coils. In addition, polymer-polymer <strong>and</strong><br />

polymer-solvent interactions exert an important effect in polymer chain configurations <strong>and</strong><br />

must be taken into account. If segments <strong>of</strong> the polymer chains stick together, then a tighter<br />

configuration than r<strong>and</strong>om coil will result, enabling aggregate formation or precipitation (3).<br />

1.2.- Classification <strong>of</strong> polymers.<br />

Polymers can be classified in many different ways. The most obvious classification is based on<br />

the origin <strong>of</strong> the polymer; i.e., natural polymers (biological macromolecules) such as proteins,<br />

celluloses or natural gums, <strong>and</strong> synthetic polymers (non biological polymers). Overall, non-<br />

biological polymers are formed by flexible chains with a relatively “low” number <strong>of</strong> monomers<br />

16


in contrast to biological polymers, which consist <strong>of</strong> a larger number <strong>of</strong> monomers <strong>and</strong> with<br />

stiffer chains.<br />

1.2.1.- <strong>Synthetic</strong> polymers.<br />

<strong>Synthetic</strong> polymers can be classified taking into account their physical <strong>and</strong> chemical<br />

characteristics, molecular structure, polymerisation mechanisms, chemical composition,<br />

building blocks or thermal behaviour. Among all <strong>of</strong> them, synthetic polymers have been<br />

classified into two classes according to their preparation method: i.e., condensation <strong>and</strong><br />

addition polymers. For condensation polymers, the reaction occurs between two<br />

polyfunctional molecules by eliminating a small molecule, for example, water. In contrast,<br />

addition polymers are formed in a chain reaction <strong>of</strong> monomers which have double bonds in<br />

their molecular structure.<br />

According to polymer structure, it is possible to divide synthetic polymers into three wide<br />

categories: linear, branched (non-linear) <strong>and</strong> cross-linked polymers. When the structural units<br />

are repeated in only one dimension, linear polymers are obtained. Their molecular structure is<br />

shown in Figure 1.1a. On the other h<strong>and</strong>, when the monomeric units are repeated in two or<br />

three dimensions through covalent bonds, branched, cross-linked or network polymers are<br />

obtained. In the case <strong>of</strong> branched polymers (also known as graft copolymers), single side-<br />

branched chains are connected to polymer backbones during the synthesis, <strong>and</strong> which remain<br />

singly dispersed (see Figure 1b). For cross-linked polymers, adjacent linear chains are<br />

chemically or physically connected to two or more polymer backbones serving as “bridges” to<br />

reticulate the polymeric system (see Figure 1.1c).<br />

Figure 1.1. a) Linear, b) branched, <strong>and</strong> c) cross-linked polymers.<br />

17


On the other h<strong>and</strong>, in terms <strong>of</strong> chemical composition, polymers can be classified as<br />

homopolymers <strong>and</strong> copolymers. Homopolymers are those polymers composed <strong>of</strong> only one<br />

type <strong>of</strong> structural unit, whereas copolymers posses two or more different monomers.<br />

Copolymers can be additionally classified considering the sequential <strong>and</strong> structural order <strong>of</strong><br />

their constituent building blocks (see Figure 1.2):<br />

R<strong>and</strong>om copolymer: the building segments in the backbone follow a r<strong>and</strong>om sequence.<br />

Alternate copolymer: the segments that configure the copolymer show a specific<br />

alternating sequence.<br />

Block copolymer: the polymer backbone is formed by the covalent bonding <strong>of</strong> large<br />

sequences <strong>of</strong> different homopolymers (each one forming a monomer).<br />

Graft copolymers: the copolymer molecule posses several segments grafted along the<br />

polymeric backbone.<br />

Figure 1.2. Schematic representation <strong>of</strong> different copolymers: (a) R<strong>and</strong>om copolymer (b)<br />

alternating copolymer, (c) <strong>and</strong> (d) block copolymer, (e) <strong>and</strong> (f) graft copolymer.<br />

1.2.2.- Biopolymers.<br />

Natural polymers are vital for the development, growth <strong>and</strong> perpetuation <strong>of</strong> any living<br />

organism. Under this consideration, the most important ones are nucleic acids, proteins <strong>and</strong><br />

polysaccharides. The building units <strong>of</strong> these biological polymers are nucleotides, amino acids,<br />

<strong>and</strong> carbohydrates, respectively. Among all these types, the present work is focused on the<br />

study <strong>of</strong> proteins since they are an essential constituent <strong>of</strong> all organisms <strong>and</strong> virtually<br />

participate in every process within cells.<br />

18


1.2.3.- Proteins.<br />

The most versatile macromolecules in living organisms are proteins. These play an important<br />

<strong>and</strong> crucial function in almost all biological processes, <strong>and</strong> are formed by the linkage <strong>of</strong><br />

repetitive structural units, called amino acids, through peptides bonds (4). These are covalent<br />

chemical bonds formed between two amino acids, in which the carboxyl group <strong>of</strong> one<br />

molecule reacts with the amino group <strong>of</strong> the other, as shown in Figure 1.3. Thus, polypeptide<br />

chains have two different terminal groups: on one side, the polypeptide chain posses an amino<br />

group with positive polarity <strong>and</strong>, on the other, a carboxylic group with negative polarity. By<br />

convention, the amino group has been considered the beginning <strong>of</strong> the polypeptide chain.<br />

Figure 1.3. Peptide bond formation between the amino group <strong>of</strong> one amino acid <strong>and</strong> the<br />

carboxyl group <strong>of</strong> the other.<br />

1.3.- (Bio)polymer structure.<br />

It is known from elemental organic chemistry that the physical state <strong>of</strong> a homologous series <strong>of</strong><br />

any type <strong>of</strong> chemical compound changes as their molecular size increases. From the low to the<br />

high molecular weight end <strong>of</strong> the molecular spectrum, the members <strong>of</strong> a series change<br />

progressively from the gaseous state, passing through liquids <strong>of</strong> increasing viscosity<br />

(decreasing volatility) to low melting solids <strong>and</strong>, ultimately, terminate in high-strength solids.<br />

Polymers belong to the high molecular weight end <strong>of</strong> the molecular spectrum, which provides<br />

them very interesting <strong>and</strong> useful properties such as unusual crystallinity, a wide range <strong>of</strong><br />

thermal <strong>and</strong> mechanical behavior, <strong>and</strong> high resistance to chemical degradation. All these<br />

properties are related to the chemical <strong>and</strong> structural aspects <strong>of</strong> polymers, which can be<br />

divided into three different levels:<br />

1. The monomer chemical structure (primary structure)<br />

2. The single polymer chain (secondary level)<br />

3. Aggregation <strong>of</strong> polymer chains (tertiary structure)<br />

19


Additionally, to obtain a better knowledge about the macromolecular structure, it is necessary<br />

to study the molecular forces existing between polymer chains, such as ionic <strong>and</strong> covalent<br />

bonds, dipole-dipole interactions, hydrogen bonding, induction forces or van der Waals<br />

interactions, amongst others (5).<br />

1.3.1.- Primary Structure.<br />

The primary structure refers to the atomic composition <strong>and</strong> chemical structure <strong>of</strong> the<br />

monomers, which compose the polymeric chains. The electrical <strong>and</strong> chemical properties <strong>of</strong><br />

polymeric chains are much related to the nature <strong>and</strong> chemistry <strong>of</strong> the constituent monomers<br />

<strong>and</strong> the kind <strong>of</strong> bonds established between them. Also, the macromolecular size <strong>of</strong> the<br />

polymer is responsible <strong>of</strong> their physical <strong>and</strong> mechanical properties.<br />

1.3.2.- Secondary Structure.<br />

To underst<strong>and</strong> the properties <strong>of</strong> polymers, it is necessary to develop a physical picture <strong>of</strong> what<br />

these long molecules are really like. The secondary structure is a real physical description <strong>of</strong><br />

the polymer, <strong>and</strong> describes the size <strong>and</strong> shape <strong>of</strong> a single individual macromolecule. The<br />

polymer molecular architecture can be influenced by both the nature <strong>of</strong> monomers <strong>and</strong> how<br />

they are bonded each other. In order to have a suitable description about the molecular<br />

architecture <strong>of</strong> a polymer, it is necessary to define two important terms: configuration <strong>and</strong><br />

conformation.<br />

Configuration refers to the relative position <strong>of</strong> the atoms in a polymeric chain, which is<br />

determined by the chemical bonds. The mechanical <strong>and</strong> thermal properties <strong>of</strong> polymers are a<br />

function <strong>of</strong> chain configuration; thus, the configuration <strong>of</strong> a polymer cannot be altered unless<br />

these bonds are broken. The different possible configurations <strong>of</strong> polymers which have the<br />

same stoichiometric molecular formula but different structural formulation are called isomers.<br />

There are two main geometrical configurations <strong>of</strong> polymers: cis <strong>and</strong> trans. These structures<br />

cannot be changed by physical means (e.g. rotation). The cis configuration originates when<br />

substituent groups are located at the same side <strong>of</strong> a carbon-carbon double bond. Trans<br />

configuration refers to the case when the chemical substituents are on opposite sides <strong>of</strong> the<br />

double bond (see Figure 1.4a,b). On the other h<strong>and</strong>, taking into account the position <strong>of</strong> the<br />

atoms in a polymer chain three main configurations are observed: atactic polymers, in which<br />

the substituent groups have a r<strong>and</strong>om orientation (below <strong>and</strong>/or above) along the polymeric<br />

backbone; isotactic polymers, where the substituent groups are in the same plane along the<br />

20


main polymeric chain; <strong>and</strong> sindiotactic polymers, in which the substituent groups are in an<br />

alternating opposite orientation along the polymer backbone (see Figure 1.4c-e).<br />

Figure 1.4. Schematic representation <strong>of</strong> the polymers configuration a) cis, b) trans, c) polymer<br />

atactic, d) isotactic polymer, <strong>and</strong> e) sindiotactic polymer.<br />

Molecular conformation refers to the special spatial orientation <strong>of</strong> a macromolecule. To<br />

modify the macromolecular conformation it is not necessary to break out covalent bonds, but<br />

it is only necessary to rotate the atoms around single bonds in the polymer backbone.<br />

Therefore, the conformation <strong>of</strong> a polymeric molecule can be changed, for example, by<br />

application <strong>of</strong> thermal energy or by interactions between the macromolecules <strong>and</strong> the solvent<br />

molecules. The great number <strong>of</strong> possible conformations <strong>of</strong> macromolecules depends on<br />

different factors such as polymer crystallinity, steric effects or their physical state (i.e., if the<br />

polymer is in solution or in its molten or solid state), with the constraint that macromolecules<br />

tend to adopt a minimum energy conformations state. Typical macromolecular conformations,<br />

as r<strong>and</strong>om coil, fully extended chain, folded chain or helix are depicted in Figure 1.5. The<br />

r<strong>and</strong>om coil is an asymmetric conformation characterized by a r<strong>and</strong>om orientation <strong>of</strong><br />

macromolecular segments. It is typical <strong>of</strong> polymers in solution, <strong>of</strong> molten polymers <strong>and</strong> solid<br />

amorphous polymers. All other conformations reported in Figure 1.5 are symmetric <strong>and</strong>,<br />

21


therefore, characteristic <strong>of</strong> solid crystalline polymers. Hence, when the concept <strong>of</strong><br />

macromolecular conformation is limited to a single macromolecule, it defines the polymer`s<br />

secondary structure.<br />

Figure 1.5. Secondary structure <strong>of</strong> polymers.<br />

1.3.3.- Tertiary Structure.<br />

Macromolecular conformation related to several macromolecules arranged together in any<br />

liquid or solid phase defines the polymer´s tertiary structure. Usual tertiary structures are<br />

depicted in Figure 1.6. These can be represented either by a tight arrangement <strong>of</strong> r<strong>and</strong>om coil<br />

chains (spaghetti-like structure) as in molten or amorphous solid polymer; by a fringed-micelle<br />

structure as observed in semi-crystalline solid polymers; or in regularly folded chains or super-<br />

helix tertiary structures as in the case <strong>of</strong> highly crystalline solid polymers.<br />

Figure 1.6. Tertiary structure <strong>of</strong> polymer.<br />

Any polymeric material in the solid state posses great amounts <strong>of</strong> polymeric molecules in the<br />

form <strong>of</strong> aggregates. The process <strong>of</strong> molecular aggregation can lead to the formation <strong>of</strong> either<br />

crystalline or amorphous materials, which depend on both the polymer molecular structure<br />

22


<strong>and</strong> the available configurations <strong>of</strong> the polymeric backbone´s building blocks. The ability <strong>of</strong><br />

polymer chains to establish a sufficient high number <strong>of</strong> dipole-dipole interactions, hydrogen<br />

bonding, or van der Waals interactions, amongst others, is also a key factor in promoting <strong>and</strong><br />

controlling the molecular aggregation process in spite <strong>of</strong> the energy involved in this kind <strong>of</strong><br />

intermolecular forces is very weak (0.5-10 kcal/mol) if compared with covalent bonds (50-100<br />

kcal/mol)<br />

1.4.- Molecular self-assembly.<br />

Molecular self-assembly is a spontaneous process by which molecules interact with each other<br />

to form well-ordered supramolecular structures. The self-association between molecular<br />

building blocks is mediated through non-covalent association <strong>of</strong> molecules (molecular<br />

recognition processes). In this way, extraordinarily complex assemblies are formed in various<br />

natural systems, <strong>and</strong> synthetic preformed molecular building blocks are assembled together by<br />

keeping a delicate balance between non-covalent repulsive <strong>and</strong> attractive interaction by using<br />

forces such as van der Waals <strong>and</strong> ionic interactions, metal coordination, H-bonding, π-π<br />

interactions or hydrophobic forces. This process has motivated many different studies in the<br />

fields <strong>of</strong> nanotechnology <strong>and</strong> materials science based on the design <strong>of</strong> new functional<br />

nanostructures (6)(7). The only requirement for self-assembly is that the building block should<br />

contain moieties able to establish non-covalent interactions in specific environments (6)(7).<br />

Thus, advanced self-assembled nanomaterials have already been prepared from many<br />

different types <strong>of</strong> molecules such as lipids, synthetic peptides, nucleic acids, low molecular<br />

weight surfactants, dendrons, transition metals, inorganic nanoparticles or polymers (8)(9).<br />

The latter have recently received great attention due to their increasingly scientific importance<br />

<strong>and</strong> vast utility in a wide range <strong>of</strong> technological applications by their ability to spontaneously<br />

self-organize into well-ordered aggregates, either in the solid state or in semidilute <strong>and</strong> dilute<br />

solutions (9)(10).<br />

Nowadays, the self-organization <strong>of</strong> molecular building blocks is probably one <strong>of</strong> the most<br />

prominent topics <strong>of</strong> modern sciences. Indeed, supramolecular self-assembly is the key strategy<br />

<strong>of</strong> Nature for fabricating multifunctional structures such as enzymes, ribosomes, membranes<br />

channel or transport proteins from rather simple molecules such peptides, sugars <strong>and</strong> lipids.<br />

For instance, Nature produces complex intermolecular structures with biomolecules, through<br />

the control <strong>of</strong> biomolecular primary, secondary, tertiary <strong>and</strong> quaternary structure. The<br />

23


presence <strong>of</strong> electrostatic charge in biomolecular primary structure provides natural systems<br />

with a molecular-level tool to control intramolecular conformation (secondary <strong>and</strong> tertiary<br />

structure) <strong>and</strong> intermolecular aggregation (quaternary structure), such as the association <strong>of</strong><br />

lipids to form cellular membranes, the protein folding mechanisms to construct functional<br />

peptides <strong>and</strong> proteins, or the supramolecular organization <strong>of</strong> relevant biological entities such<br />

as enzymes, ribosomes or transport channels.<br />

In order to mimic the self-assembly processes observed in Nature (10)(11), block copolymers<br />

have become <strong>of</strong> great scientific importance because <strong>of</strong> their unique <strong>and</strong> associative properties<br />

as a consequence <strong>of</strong> their molecular structure. In particular, synthetic block copolymers can<br />

self-organize spontaneously into unprecedented morphologies (which are related to their<br />

segmental incompatibility), either in solid state or in semidilute or dilute solutions. For<br />

instance, in bulk block copolymers made <strong>of</strong> thermodynamically incompatible segments phase-<br />

separate into organized discrete nanophases, which can be exploited in many applications<br />

ranging from impact-resistant materials to nanolithography <strong>and</strong> nanoelectronics. Similarly, in<br />

dilute solutions block copolymers lead to the formation <strong>of</strong> colloidal aggregates such as<br />

polymeric micelles or polymer vesicles, which have an enormous potential for applications in<br />

cosmetic, the food industry, medicine <strong>and</strong> biotechnology.<br />

In solution, the main attraction forces leading to block copolymer self-assembly are:<br />

hydrophobic <strong>and</strong> electrostatic interactions, <strong>and</strong> hydrogen bonding. In this context, these<br />

macromolecules can be considered as building blocks. In a selective solvent, blocks that can be<br />

solubilized in this solvent tend to get in contact with the main phase, whereas blocks that are<br />

insoluble in the solvent tend to be hidden from the main phase, leading to the formation <strong>of</strong><br />

organizated aggregates such spherical micelles, cylindrical micelles or vesicles (12). However,<br />

amphiphilicity is not the only driving force that can be exploited for creating aggregates in<br />

solution. The pairing <strong>of</strong> complementary charges in the polymer backbone is also a vector for<br />

self-organization. Hydrogen bonding interactions are also important for the development <strong>of</strong><br />

self-assembling supramolecular aggregates, in which monomeric units are reversibly bound via<br />

weak non-covalent interactions to form well-ordered structures.<br />

For example, due to the great current interest in the development <strong>and</strong> improvement <strong>of</strong> nano-<br />

devices, many <strong>of</strong> these synthetic block copolymers are being used in nanopatterning using the<br />

top-down <strong>and</strong> bottom-up fabrication methods (Figure 1.7). The top-up strategy uses physical<br />

engineering tools for carving <strong>of</strong> materials with a considerable macro-size to obtain a small<br />

24


nanodevice. During the process, the lateral dimension <strong>of</strong> the material is reduced until having a<br />

nanostructured material as occurred, for example, in the case <strong>of</strong> silicon integrated circuits<br />

fabricated by selective layer deposition. In contrast, the bottom-up strategy takes advantage<br />

<strong>of</strong> the self-association properties <strong>of</strong> molecular blocks, which possess nanometric size. The<br />

living cell is a complex construction given by the Nature, <strong>and</strong> it is the best example for the<br />

bottom-up strategy (8)(13)(14).<br />

Figure 1.7. Schematic representation <strong>of</strong> two both strategies used to get nanostructured<br />

materials: a) Top-down <strong>and</strong> b) Bottom-up.<br />

Top-down<br />

Physical engineering<br />

Bottom-up<br />

Supramolecular self-assembling<br />

Nanostructural<br />

Material<br />

On the other h<strong>and</strong>, peptides <strong>and</strong> proteins show also amphipathic properties. The<br />

intramolecular fold <strong>of</strong> these macromolecules allows to show specific faces or display specific<br />

regions toward the bulk solvent phase; in this way, it is logical to assume that these regions<br />

have hydrophilic or hydrophobic nature. For instance, the cylindrical structure <strong>of</strong> the α-helix<br />

contain a sequence <strong>of</strong> hydrophobic amino acids which are positioned on one side <strong>of</strong> the<br />

cylindrical face, <strong>and</strong> another sequence <strong>of</strong> hydrophobic amino acids on the opposite face <strong>of</strong> the<br />

cylindrical structure. For β-sheet structure, hydrophobic <strong>and</strong> hydrophilic amino acids <strong>of</strong> a<br />

polypeptide chain can be ordered in an alternating way. In this manner, the side lateral chains<br />

<strong>of</strong> the amino acids are on opposite faces into the β-sheet structure (8).<br />

25


Contents<br />

2.1<br />

2.2<br />

2.3<br />

2.4<br />

2.5<br />

2.6<br />

2.7<br />

2.8<br />

2.9<br />

2.10<br />

2.11<br />

2.12<br />

2.2.1<br />

2.2.2<br />

2.2.3<br />

2.3.1<br />

2.3.2<br />

2.4.1<br />

2.4.2<br />

2.9.1<br />

2.11.1<br />

2.11.2<br />

2.11.3<br />

2.11.4<br />

2.11.5<br />

Introduction<br />

Surface tension<br />

Wilhelmy plate method<br />

Monolayers <strong>and</strong> Langmuir-Blodgett films<br />

Axisymetric drop shape (ADSA)<br />

Light scattering<br />

Static light scattering (SLS)<br />

Dynamic light scattering (DLS)<br />

Electron microscopy<br />

Transmission electron microscopy (TEM)<br />

Scanning electron microscopy (SEM)<br />

Atomic force microscopy (AFM)<br />

Fluorescence spectroscopy<br />

Circular dichroism (CD)<br />

Infrared spectroscopy<br />

Rheology<br />

Viscoelasticity<br />

X-ray diffraction<br />

CHAPTER 2<br />

Methods<br />

Additional experimental techniques used in fibrillogenesis<br />

studies<br />

UV-Vis spectroscopy<br />

Polarized light optical microscopy<br />

Emission spectroscopy<br />

Superconducting quantum interference device (SQUID)<br />

Magnetic resonance imaging (MRI)<br />

Laboratory equipment<br />

28<br />

28<br />

28<br />

29<br />

31<br />

32<br />

36<br />

40<br />

44<br />

44<br />

46<br />

47<br />

48<br />

52<br />

54<br />

58<br />

59<br />

64<br />

72<br />

72<br />

75<br />

77<br />

78<br />

80<br />

82<br />

27


2.1.- Introduction<br />

There are many analytical techniques suitable for the characterization <strong>of</strong> synthetic <strong>and</strong><br />

biological macromolecules. This chapter is a brief summary <strong>of</strong> the basic theoretical background<br />

<strong>and</strong> experimental setups used to perform the different experiments discussed in this Thesis.<br />

In aqueous solutions <strong>of</strong> amphiphilic self-assembled systems, the aggregation process mostly<br />

depends on concentration or/<strong>and</strong> temperature. The most usual experimental tools to follow<br />

their self-assembly process into structural aggregates (micelles, vesicles...) are surface tension,<br />

fluorescence spectroscopy, light scattering or nuclear magnetic resonance, amongst others. On<br />

the other h<strong>and</strong>, the structural characteristics <strong>of</strong> the resulting self-assembled nanostructures<br />

can be determined by techniques such as X-ray diffraction, light scattering or electron <strong>and</strong><br />

atomic microscopies.<br />

2.2.- Surface tension<br />

The critical micelle concentration (cmc) <strong>of</strong> amphiphilic systems <strong>and</strong>, in particular, in block<br />

copolymer solutions can be determined by measuring their surface tension () as a function <strong>of</strong><br />

concentration. Surface tension depends on polymer concentration below its cmc <strong>and</strong>,<br />

therefore, such behavior is denoted by a fairly sharp decrease in a vs. log (c) plot. There are<br />

several methods for determining the surface tension <strong>of</strong> amphiphilic polymer solutions such as<br />

the Du Noüy ring, the Wilhelmy plate, the pendant drop, the bubble pressure or the sessile<br />

drop methods. The most common ones are the Du Noüy ring <strong>and</strong> Wilhelmy plate techniques<br />

(15).<br />

2.2.1.- Wilhelmy plate method<br />

The Whilhelmy plate method measures the force (F) with which a platinum plate <strong>of</strong> known<br />

perimeter (L= l+d) is pulled downward by an interface, in our particular case the air/water<br />

<strong>and</strong> air/chlor<strong>of</strong>orm interfaces. The surface tension force, Lcos , is equal to the weight <strong>of</strong> the<br />

liquid meniscus adsorbed onto the plate <strong>and</strong> detected by a balance (16), which can be<br />

expressed through the following equation:<br />

2.01<br />

28


where w is the meniscus’ weight detected by the balance, is the contact angle defined by the<br />

meniscus shape on the wet platinum surface plate, <strong>and</strong> l is the width <strong>and</strong> d the plate thickness,<br />

respectively (Figure 2.1). Considering very small <strong>and</strong> the plate thickness negligible compared<br />

with its width, the former expression is simplified to:<br />

Figure 2.1. Illustration <strong>of</strong> a fully wet Wilhelmy plate in contact with an aqueous solution.<br />

2.2.2.- Monolayers <strong>and</strong> Langmuir-Blodgett films<br />

Many insoluble substances such as long chain fatty acids, alcohols, surfactants <strong>and</strong> polymers<br />

can be spread from a solvent onto water to form a one-molecule thick film, called monolayer.<br />

Molecules in a monolayer can be arranged in different ways, especially when they are closely<br />

packed, depending on the lateral forces between them (17). The obtention <strong>of</strong> surface<br />

pressure-area isotherms is one <strong>of</strong> the main methods used to study monolayer formation.<br />

Surface pressure, , is the reduction <strong>of</strong> the surface tension due to the presence <strong>of</strong> the<br />

monolayer: , where 0 is the surface tension <strong>of</strong> the solvent (usually water), <strong>and</strong> <br />

the surface tension <strong>of</strong> the system. Surface pressure-area isotherms are <strong>of</strong>ten measured for<br />

films/monolayers compressed using a Langmuir trough (Figure 2.2).<br />

In general, an insoluble substance is spread from a volatile solvent to ensure the formation <strong>of</strong> a<br />

uniform monolayer, <strong>and</strong> its area is controlled through a moveable barrier. The horizontal force<br />

necessary to maintain the float at a fixed position is measured using a torsion balance, which<br />

provides the surface tension values. Since only micrograms <strong>of</strong> materials are present in<br />

2.02<br />

29


monolayers, it is necessary to ensure that the water is as pure as possible (17). On the other<br />

h<strong>and</strong>, it is also possible to deposit the monolayer film onto a solid substrate, or build up a<br />

multilayer by successive depositions <strong>of</strong> monolayer films. In this case, this technique is called<br />

Langmuir-Blodgett <strong>and</strong> the films are named Langmuir-Blodgett (or LB) films. A monolayer is<br />

adsorbed homogeneously with each immersion or emersion step, thus films with very accurate<br />

thickness can be formed. This thickness is accurate because the thickness <strong>of</strong> each monolayer is<br />

known <strong>and</strong> can therefore be added to find the total thickness <strong>of</strong> a Langmuir-Blodgett film.<br />

These LB films are <strong>of</strong> great interest since they enable the construction <strong>of</strong> functional multilayer<br />

stacks which can be used in molecular electronic devices, such as gas sensors, diodes or in<br />

semiconductor devices as insulators in transistors.<br />

Figure 2.2. Scheme <strong>of</strong> a Langmuir trough.<br />

When a monolayer is compressed on the water surface, several phase transformations can be<br />

observed. These changes are almost analogous to those present in bulk phase matter, where<br />

the most common states are gas, liquid <strong>and</strong> solid. The phase changes may be readily identified<br />

by monitoring the surface pressure as a function <strong>of</strong> the area occupied by the film (surface<br />

pressure-area isotherm) (18). This is the two dimensional equivalent <strong>of</strong> the pressure-volume<br />

isotherm for bulk materials (gas/liquid/solid). Figure 2.3 shows the surface pressure-area<br />

isotherm plot <strong>of</strong> a hypothetical long chain organic monolayer material (e.g., a long-chain fatty<br />

acid). This diagram displays most <strong>of</strong> the characteristic features observed for long chain<br />

compounds (18). The shape <strong>of</strong> the surface pressure-area isotherm depends on the lateral<br />

interactions between molecules. These, in turn, depends on the molecular packing, which is<br />

influenced by factors such as the size <strong>of</strong> the hydrophilic group, the presence <strong>of</strong> polar groups,<br />

the length <strong>and</strong> number <strong>of</strong> hydrocarbon chains, <strong>and</strong> their conformation.<br />

Briefly, in the gaseous state (G) the molecules are far enough apart on the water surface so<br />

30


that they exert little force one on another. As the surface area <strong>of</strong> the monolayer is reduced,<br />

the hydrophobic chains <strong>of</strong> the molecules will begin to interact. The liquid state formed is<br />

generally called the exp<strong>and</strong>ed monolayer phase (E). At this stage, molecules at the water<br />

surface are in a r<strong>and</strong>om rather than in a regular orientation, with their polar groups in contact<br />

with the subphase. As the molecular area is progressively reduced, a condensed (C) phase may<br />

appear. In the condensed monolayer state the molecules are closely packed <strong>and</strong> are oriented<br />

with the hydrophobic groups pointing away from the water surface.<br />

Figure 2.3. Surface pressure-area isotherm for a chain organic monolayer.<br />

2.2.3.- Axisymetric drop shape analysis (ADSA)<br />

Axisymetric drop shape analysis (ADSA) is a drop shape technique for measuring the interfacial<br />

surface tension, γ. This method can be performed on the measured pr<strong>of</strong>iles <strong>of</strong> captive bubbles,<br />

sessile drops, pendant drops or bubbles. The principle <strong>of</strong> the technique is based on the<br />

determination <strong>of</strong> the drop shape by the balance between: i) the force <strong>of</strong> gravity which deforms<br />

the drop <strong>and</strong> ii) the surface tension which resists the deformation by minimizing the drop<br />

surface area (17). The relationship between these two forces is described by the Laplace<br />

equation (19):<br />

The Laplace equation is written in terms <strong>of</strong> the surface tension, , the principal radii <strong>of</strong> the<br />

drop, <strong>and</strong> , <strong>and</strong> the pressure difference across the curved interface, . The pressure<br />

drop has two contributions, , where is the pressure difference at a<br />

2.03<br />

31


eference plane, <strong>and</strong> the term gz is an additional pressure which arises from the gravitational<br />

component at any point on the surface. This is given by the surface density difference, ,<br />

between the bubble <strong>and</strong> the continuous phase, <strong>and</strong> its vertical distance from the reference<br />

plane (17).<br />

ADSA technique allows to determine both the surface area, , as well as the surface tension, .<br />

Additionally, both quantities enable the derivation <strong>of</strong> important surface rheological properties,<br />

such the elasticity or the surface dilatational modulus, E, <strong>of</strong> a film, defined as<br />

. This can also be defined as a complex function written as<br />

, where E’ <strong>and</strong> E’’ correspond to the storage <strong>and</strong> loss moduli, respectively, by analogy<br />

with 3D rheology. Assuming the mechanical properties <strong>of</strong> the adsorbed layer follow Maxwell’s<br />

behavior, the storage <strong>and</strong> loss moduli can be written as (20)-(22)<br />

2.04a<br />

2.04b<br />

where E0i <strong>and</strong> 0i ( are the corresponding weights <strong>of</strong> the i-th relaxation element<br />

to the total elasticity modulus, E0, <strong>and</strong> the characteristic relaxation frequency, respectively,<br />

<strong>and</strong> 0i is the characteristic relaxation time <strong>of</strong> the Maxwell model.<br />

2.3.- Light scattering<br />

Scattering is a general physical process in which a form <strong>of</strong> radiation, such as, sound or a<br />

moving particle, are forced to deviate from its straight trajectory by one or more localized non-<br />

uniformities in the medium through it propagates. In the case <strong>of</strong> light, it is fairly simple to<br />

underst<strong>and</strong> the origin <strong>of</strong> its scattering considering it as an electromagnetic wave. Light will<br />

interact with the charges inside a given molecule remodelling their spatial charge distribution.<br />

The quantification <strong>of</strong> this effect is reported by a certain physical quantity, the polarizability (α).<br />

The charge distribution follows the time-modulation <strong>of</strong> the electric wave vector <strong>of</strong> the incident<br />

light beam <strong>and</strong>, therefore, the molecule constitutes an oscillating dipole or electric oscillator.<br />

32


This oscillating dipole acts as an emitter <strong>of</strong> an electromagnetic wave with the same wavelength<br />

as the incident one (the scattering process is elastic), which is emitted isotropically in all<br />

perpendicular directions to the oscillator, as illustrated in Figure 2.4. The angle <strong>of</strong> observation<br />

with respect to the direction <strong>of</strong> the incident light beam is called the scattering angle, <strong>and</strong> it<br />

provides a measure <strong>of</strong> the accessible length scales by a light scattering experiment (23).<br />

Figure 2.4. Oscillating dipole induced by an incident light wave, <strong>and</strong> emitting light (23).<br />

For molecules or particles equal or larger than /20, being the wavelength <strong>of</strong> the incident<br />

radiation, several <strong>of</strong> these oscillating dipoles are created simultaneously within one given<br />

particle. As a consequence, some <strong>of</strong> the emitted light waves possess a significant phase<br />

difference. Accordingly, interference <strong>of</strong> the scattered light emitted from such individual<br />

particles leads to a non-isotropic angular dependence <strong>of</strong> the scattered light intensity. The<br />

interference pattern <strong>of</strong> intra-particular scattered light, also called the particle form factor, is<br />

characteristic <strong>of</strong> the geometry <strong>of</strong> the scattering particle. Hence, this provides a quantitative<br />

means for the structural characterization <strong>of</strong> particles in very dilute solutions by light scattering.<br />

On the other h<strong>and</strong>, for particles smaller than /20, only a negligible phase difference exists<br />

between light emitted from various scattering centers within a given particle; in this case, the<br />

detected scattered intensity will be independent <strong>of</strong> the scattering angle, <strong>and</strong> it only will<br />

depend on the particle mass, which is proportional to the total number <strong>of</strong> scattering centers<br />

one particle contains. The difference in the interference pattern <strong>of</strong> light scattered by small <strong>and</strong><br />

large particles, is illustrated in Figure 2.5 (23).<br />

So far, we have considered light scattering as a purely elastic process where the emitted light<br />

has exactly the same wavelength as the incident light. Particles in solution, however, usually<br />

show a r<strong>and</strong>om motion (Brownian motion) caused by thermal density fluctuations <strong>of</strong> the<br />

solvent. As a consequence <strong>of</strong> temporal changes in inter-particle positions <strong>and</strong> the<br />

corresponding temporal concentration fluctuations, both the interference pattern <strong>and</strong> the<br />

resulting scattered intensity detected at a given scattering angle also change with time,<br />

33


eflecting the Brownian motion <strong>of</strong> the scattering particles (23). This phenomenon provides the<br />

basis <strong>of</strong> dynamic light scattering, an experimental procedure which yields a quantitative<br />

measure for the mobility <strong>of</strong> scattering particles in solution, characterized by their self-diffusion<br />

coefficients. Most modern particle size analyzers used to determine the size <strong>of</strong> particles in<br />

solution are based on this principle, further commented in detail below (23)-(25).<br />

Figure 2.5. Interference pattern <strong>of</strong> light scattered from small particles (left) <strong>and</strong> large particles<br />

(right). For simplification, only two scattering centers are shown (23).<br />

General st<strong>and</strong>ard light scattering setups consist <strong>of</strong> the following components (Figure 2.6) (23)-<br />

(26):<br />

1. The incident light source, typically a laser source. This provides coherent <strong>and</strong><br />

monochromatic light with an intensity between some few milliwatts (mW) <strong>and</strong> several watts<br />

(W). In practice, the light intensity needed for a successful scattering experiment depends on<br />

the sensitivity <strong>of</strong> the optical detector <strong>and</strong> on the scattering power <strong>of</strong> the sample itself, as<br />

determined by its size, concentration, <strong>and</strong> refractive index increment. As shown in Figure 2.6,<br />

typically, the primary laser beam is guided <strong>and</strong> focused onto the sample by optical mirrors <strong>and</strong><br />

lenses. The laser beam diameter within the sample is well below 1 mm, which defines the<br />

scattering volume.<br />

2. The light scattering cell, in most cases a cylindrical quartz glass cuvette <strong>of</strong> outer<br />

diameter between 10 <strong>and</strong> 30 mm, embedded within an index matching <strong>and</strong> thermostating<br />

bath. Using an index matching bath around the cylindrical light scattering cuvette is important<br />

to suppress unwanted diffraction effects <strong>of</strong> the incident <strong>and</strong> the scattered light at the sample-<br />

air interfaces.<br />

3. The detector, either a photo-multiplier tube or a very sensitive avalanche photo diode<br />

34


(APD) <strong>and</strong> its associated optics (pinhole or optical fiber) mounted on the arm <strong>of</strong> a goniometer.<br />

The detector optics determines the horizontal <strong>and</strong> vertical dimensions <strong>of</strong> the scattering<br />

volume, whereas its depth is defined by the incident laser beam width. B<strong>and</strong> pass filters with<br />

high transmission at the incident laser light wavelength are <strong>of</strong>ten used in front <strong>of</strong> the detector<br />

to suppress undesired contributions <strong>of</strong> stray light or fluorescence from the sample.<br />

4. The electronic hardware components associated with the detector used for signal<br />

processing. The scattered intensity I (q,t), which is monitored by the optical detector, is<br />

typically digitalized <strong>and</strong> stored in a computer. These raw data have to be further processed: in<br />

the case <strong>of</strong> static light scattering, particle form factor models can be used to fit the<br />

experimental data; in a dynamic light scattering experiment, the raw data have to be<br />

converted into the time intensity correlation function g2 (q, τ). This is done during the<br />

measurement using a very fast hardware correlator as, for example, the ALV 5000 from ALV,<br />

Langen, Germany.<br />

5. In a typical light scattering experiment, light from a laser source passes through a<br />

polarizer, which defines the polarization <strong>of</strong> the incident beam <strong>and</strong>, then, enters the scattering<br />

medium. The scattered light, then, passes through an analyser, which selects a given<br />

polarization <strong>and</strong> finally enters a detector. The position <strong>of</strong> the detector defines the scattering<br />

angle, θ. In addition, the intersection <strong>of</strong> the incident beam <strong>and</strong> the beam intercepted by the<br />

detector defines a scattering region <strong>of</strong> volume V. In modern light scattering experiments the<br />

scattered light spectral distribution is measured by a photomultiplier, which is the main<br />

detector (24).<br />

Figure 2.6. Schematic representation <strong>of</strong> an usual light scattering experiment (26).<br />

35


2.3.1.- Static light scattering (SLS)<br />

As mentioned above, matter scatters electromagnetic waves due to the induction <strong>of</strong> an<br />

oscillating electromagnetic dipole which serves as a source for the scattered light wave. In this<br />

way, Rayleigh scattering is described in terms <strong>of</strong> three factors: the incident light, the particle<br />

(i.e. a macromolecule), which serves as an oscillating dipole, <strong>and</strong> the scattered light. The<br />

schematic model is shown in Figure 2.7. The equation for the electric field <strong>of</strong> incident polarized<br />

light beam at x position <strong>and</strong> at time, t, may be expressed by the following well-known equation<br />

(27)-(32):<br />

where E is the electric field or electric intensity, E0 is the amplitude <strong>of</strong> the incident electrical<br />

wave, ν is the light frequency in solution, <strong>and</strong> t is the propagation time.<br />

Figure 2.7. Sketch <strong>of</strong> a light scattering process (23).<br />

If an incident light reach a particle (i.e. macromolecule), whose size is smaller than the<br />

wavelength , d


where r is the distance <strong>of</strong> the dipole from the observer (from the scattered light sample to the<br />

detector) <strong>and</strong> is the angle between the dipole axis <strong>and</strong> the line r. The division by the square<br />

<strong>of</strong> the light velocity, c 2 , is a dimensional correction. Substituting<br />

equation:<br />

2.07<br />

in the above<br />

Comparing the intensity I <strong>of</strong> the scattered radiation to the intensity I0 <strong>of</strong> the incident radiation,<br />

which is proportional to the square <strong>of</strong> its amplitude E0:<br />

This equation is called the Rayleigth equation for plane polarized light. The scattering intensity<br />

increases rapidly as the light wavelength decreases. The intensity also depends on the angle :<br />

there is no radiation along the direction ( =0) in which the dipole vibrates.<br />

Let us consider a system <strong>of</strong> volume V containing N independent identical scattering particles.<br />

Then, the total intensity I <strong>of</strong> scattered light is:<br />

Generally, the Rayleight scattering equation is directly applicable to gases, where molecules<br />

move r<strong>and</strong>only. In liquids, there exists fluctuations in particle concentration inside a volume<br />

element, which results in fluctuations <strong>of</strong> the polarizability α. Therefore, to apply the Rayleigh<br />

scattering equation to the liquid state, it is necessary to account for these fluctuations by using<br />

α. In this way, since the polarizability α depens on the dielectric permitittivity ε, <strong>and</strong><br />

correspondingly on the index <strong>of</strong> refraction n, then:<br />

2.08<br />

2.09<br />

2.10<br />

2.11<br />

37


For a dilute solution, polarizability α may be written as:<br />

where dn/dc is the differential refractive index <strong>and</strong> is an experimentally measurable quantity.<br />

Substitution <strong>of</strong> the equation 2.12 into equation 2.10, <strong>and</strong> considering that<br />

expressed as<br />

number:<br />

Defining the Rayleigh ratio <strong>and</strong> is <strong>of</strong>ten designed as R:<br />

which is independents <strong>of</strong> r <strong>and</strong> . Simplifying:<br />

Then the light scattering becomes:<br />

where<br />

2.12<br />

can be<br />

, where M is the molecular weight <strong>and</strong> NA is the Avogadro’s<br />

.<br />

As mentioned above, for very dilute solution <strong>of</strong> small particles (for example, nanoparticles or<br />

polymer chains <strong>of</strong> size smaller than /20) the scattering intensity is independent <strong>of</strong> the<br />

scattering angle, <strong>and</strong> the density fluctuation from the surrounding solvent can be substracted<br />

by considering the excess Rayleigh ratio, R ( ; then, the scattering<br />

intensity only depends on the scattering power <strong>of</strong> the dissolved particles (i.e., from their<br />

molecular mass) <strong>and</strong> the osmotic pressure. For non ideal solutions, <strong>and</strong> after several<br />

theoretical considerations (3), the scattering equation can be rewritten:<br />

2.13<br />

2.14<br />

2.15<br />

2.16<br />

38


When , the former equation can be simplified to the so-called Rayleigh-Gans-Debye<br />

relation:<br />

Actually, ΔR cannot be experimentally measured since I0 <strong>and</strong> r are unknown during the<br />

experiment. However, in a routine experiment ΔR is determined from the experimentally<br />

measured scattered intensities <strong>of</strong> the solution, Isolution, <strong>and</strong> a reference substance, Iref (usually<br />

benzene or toluene) (23) which enables the derivation <strong>of</strong> their respective Rayleigh ratios.<br />

Hence, the following equation can be derived providing a relation between the molecular<br />

weight, Mw, <strong>of</strong> the particle considered <strong>and</strong> the second virial coefficient, A2 (29)-(30):<br />

where<br />

,<br />

<strong>and</strong><br />

2.17<br />

2.18<br />

2.19<br />

. I, Is <strong>and</strong> Iref, are<br />

the experimentally measured scattered intensities <strong>of</strong> the solution, solvent, <strong>and</strong> reference,<br />

respectively. For small particles at dilute concentrations equation 2.19 can be reduced to:<br />

On the other h<strong>and</strong>, for large scattering particles the scattered intensity is no longer<br />

independent <strong>of</strong> the scattering angle. The scattering vector q, which is experimentally<br />

determined by the scattering angle <strong>and</strong> the laser light wavelength, , provides a quantitative<br />

measure <strong>of</strong> the length scale <strong>of</strong> the static light scattering experiment, given by:<br />

For very dilute solutions, interference between different scattering particles, the so-called<br />

2.20<br />

2.21<br />

39


structure factor, can be neglected. In this case, the angular dependence <strong>of</strong> the measured<br />

scattered intensity I(q) is only caused by intraparticle interference. For dilute <strong>and</strong> semidilute<br />

solutions, to take into account the effects <strong>of</strong> particle concentration <strong>and</strong> solute-solvent<br />

interactions on the measured scattering intensity, the Zimm equation can be used:<br />

In order to determine Mw, Rg, <strong>and</strong> A2 it is necessary to prepare several dilute solutions <strong>of</strong><br />

different concentration <strong>and</strong> measuring the scattering intensities data at various scattering<br />

angles (27)(28).<br />

2.3.2.- Dynamic light scattering (DLS)<br />

Light scattered by a set <strong>of</strong> mesoscopic particles produces a r<strong>and</strong>om interference pattern on a<br />

screen. This pattern, generally possess the form <strong>of</strong> irregularly spaced <strong>and</strong> sized bright spots<br />

called speckles. The scattered light intensity pattern remains unchanged as long as the<br />

particles do not change their position. Particle motion naturally leads to a temporal evolution<br />

<strong>of</strong> the scattered speckle field since one interference pattern is continuously replaced by<br />

another. In a single speckle spot, this evolution is observed as strong temporal intensity<br />

fluctuations, with a certain well-defined temporal correlation. The intensity fluctuations are<br />

inherently related to the scatterers’ dynamics <strong>and</strong>, therefore, the temporal correlation<br />

function depends on the particles shift. Thus, the measurable correlation properties <strong>of</strong> light<br />

can be linked to the dynamical properties <strong>of</strong> particles which, in turn, can provide information<br />

about their flow velocity <strong>and</strong> direction, particle size, density <strong>of</strong> moving scatterers...<br />

In dynamic light scattering, also known as photon correlation spectroscopy (PCS) or quasi-<br />

elastic light scattering (QELS), the diffusive motions <strong>of</strong> particles in solution gives rise to<br />

fluctuations in the scattered light intensity on the microsecond timescale. This technique is<br />

one <strong>of</strong> the most popular methods used to determine the particles size by measuring the<br />

temporal fluctuations <strong>of</strong> the scattered light intensity. Roughly, this is made by focusing a<br />

monochromatic light beam, such a laser, on a solution with particles in Brownian motion; this<br />

causes a Doppler shift when the light “hits” the moving particle, changing the light wavelength.<br />

This change is related to the size particle (27)(32).<br />

In the scope <strong>of</strong> DLS, temporal fluctuations are usually analyzed by means <strong>of</strong> the intensity<br />

2.22<br />

40


autocorrelation function (ACF). In the time domain, the correlation function usually decays<br />

with time (Figure 2.8), <strong>and</strong> faster dynamics leads to faster decorrelation <strong>of</strong> the scattered<br />

intensity trace. It can be shown that for a r<strong>and</strong>om process the intensity ACF is the Fourier<br />

transform <strong>of</strong> the power spectrum <strong>and</strong>, therefore, DLS measurements can be equally well-<br />

performed in the spectral domain. In fact, DLS experiments were initially discussed in terms <strong>of</strong><br />

the broadening <strong>of</strong> the spectrum peak <strong>of</strong> monochromatic light due to Doppler shifts<br />

experienced by propagating light waves scattered by moving particles (31).<br />

Figure 2.8. Illustration <strong>of</strong> autocorrelation function (31).<br />

To detect the intensity fluctuation with time, a DLS system requires an autocorrelator on top<br />

<strong>of</strong> a regular SLS system, as shown in Figure 2.9.<br />

Figure 2.9. Sketch <strong>of</strong> a dynamic light scattering system. The pulse amplifier discriminator<br />

converts the analogic signal <strong>of</strong> the photodetector, I(t), in a digitalized signal, which is further<br />

processed by the autocorrelator into the autocorrelation function.<br />

41


As commented above, Figure 2.10a illustrates how the intensity (I) varies with time (t). I(t)<br />

fluctuates around its mean value, . It may appear completely r<strong>and</strong>om <strong>and</strong>, therefore,<br />

meaningless, but it is not. Motions <strong>of</strong> particles (i.e. polymer molecules) <strong>and</strong> solvent molecules<br />

contribute to the change <strong>of</strong> I(t) with time. This apparently noisy signal carries the information<br />

about particles motions.<br />

Figure 2.10. a) Light scattering intensity I(t) fluctuates around its mean value . b)<br />

Autocorrelation function is obtained as the long-time average <strong>of</strong><br />

for various delay times . The aurocorrelation function decays from to over time. The<br />

amplitude <strong>of</strong> the decaying component is .<br />

The autocorrelator calculates the product average <strong>of</strong> two scattering intensities I(t) <strong>and</strong> I(t+)<br />

measured at two different times separated by , the delay time. The average product<br />

is a function <strong>of</strong> <strong>and</strong> is called the correlation function <strong>of</strong> I(t), or the intensity-<br />

intensity correlation function. The correlator converts I(t) into . Hence, the<br />

intensity-intensity correlation function is the average <strong>of</strong> over a long period T.<br />

Hence, assuming that the long time average is equal to the ensemble average, we can write:<br />

Note that, if the system is at equilibrium, the ensemble average does not change with time<br />

2.23<br />

42


<strong>and</strong>, therefore, . The autocorrelation function <strong>of</strong> I(t) (Figure 2.10a)<br />

is shown in Figure 2.10b. When = 0, . With increasing ,<br />

decays to an asymptotic level (baseline), .<br />

Because the scattering intensity I(t) fluctuates around a mean value , it is convenient to<br />

separate its fluctuating component, , as . By definition, .<br />

Then, the correlation function can be rewritten as:<br />

The autocorrelation <strong>of</strong> is zero upon increases <strong>of</strong> . When ,<br />

2.24<br />

. The decaying componet in is . The<br />

initial height <strong>of</strong> the decaying component is . Division <strong>of</strong> by<br />

leads to the intensity autocorrelation function as:<br />

where<br />

normalized intensity autocorrelation function:<br />

2.25<br />

(the coherent factor), <strong>and</strong> the second factor is the baseline-subtracted,<br />

The coherent factor depends on the coherence <strong>of</strong> the light reaching the photodetector. The<br />

measured intensity correlation function is related to the field correlation function by the<br />

Siegert relation:<br />

where is:<br />

where is the electric field conjugate function. For monodisperse spherical particles:<br />

2.26<br />

2.27<br />

2.28<br />

43


where is the characteristic delay rate, which is related to the translational diffusion<br />

coefficient, D, <strong>of</strong> a solute by means <strong>of</strong> the expression:<br />

The diffusion coefficient is frequently used to determine the hydrodynamic radius, RH, <strong>of</strong> the<br />

constituent particles by using the Stokes-Einstein equation:<br />

where KB is the Boltzmann constant, T the absolute temperature <strong>and</strong> the liquid viscosity.<br />

The hydrodynamic radius obtained by DLS represents an ideal hard sphere that diffuses with<br />

the same speed as the particle under examination. Actually, particles are solvated <strong>and</strong> the<br />

radius calculated from the particle diffusion corresponds to the size <strong>of</strong> the dynamic solvated<br />

particle.<br />

2.4.- Electron microscopy<br />

Microscopy is an important tool for the characterization <strong>of</strong> nanomaterials structure. For<br />

example, this set <strong>of</strong> imaging techniques may be used to examine the detailed shape, size <strong>and</strong><br />

distribution <strong>of</strong> a wide range <strong>of</strong> nanoscopic objects. There are different available classes <strong>of</strong><br />

microscopic techniques, such as the traditional optical microscopy, or more powerful ones<br />

such as scanning electron microscopy (SEM), transmission electron microscopy (TEM), <strong>and</strong><br />

scanning probe microscopy (SPM), amongst others.<br />

2.4.1.- Transmission electron microscopy (TEM)<br />

Transmission electron microscopy (TEM) involves the transmission <strong>of</strong> an electron beam<br />

through a sample in a high-vacuum environment. The images <strong>and</strong> their associated contrasts<br />

arise from regional differences in electron densities in such sample. TEM has a resolution <strong>of</strong> ca.<br />

1 to 100 nm <strong>and</strong> can provide very detailed structural information <strong>of</strong> polymeric materials. The<br />

TEM specimens need to be very thin in order to enable the transmission <strong>of</strong> electron beams<br />

through the sample.<br />

2.29<br />

2.30<br />

2.31<br />

44


A transmission electron microscope (Figure 2.11a) is constituted by an electron source<br />

(electron gun) which emits electrons travelling through the vacuum created into the<br />

microscope column. A condenser lens focus the electron beam on the sample, <strong>and</strong> several<br />

objective lenses are used to form the diffraction pattern at the back focal plane <strong>and</strong> the<br />

sample image at the image plane; there are also some intermediate lenses to magnify the<br />

image or the diffraction screen. To obtain an improved contrasted image, an objective<br />

diaphragm is inserted in the back focal plane to select the transmitted beam. At the bottom <strong>of</strong><br />

the microscope, the non-scattered electrons reach a fluorescent screen, which provides a<br />

contrasted image <strong>of</strong> the specimen, with a darkness distribution corresponding to its different<br />

electron densities.<br />

Figure 2.11. Schematic diagram <strong>of</strong> electron microscopes: a) TEM; b) SEM .<br />

In order to solve the material´s structure on the mesoscopic scale at even high resolutions<br />

comparable to interatomic distances, high-resolution transmission electron microscopes<br />

(HRTEM) emerge as a powerful tool. HRTEM uses the transmitted <strong>and</strong> scattered electron<br />

beams directed to a sample to create a diffraction image, in contrast to conventional TEM that<br />

only uses the transmitted beams. The high resolution image recorded can be interpretable<br />

directly in terms <strong>of</strong> projections <strong>of</strong> individual atomic positions. In this way, defects <strong>and</strong> other<br />

inhomogeneities present in a material specimen can be accurately determined. A great<br />

advantage <strong>of</strong> the HRTEM microscope is the capability to observe both electron microscope<br />

images <strong>and</strong> diffraction patterns for determined selected area electron diffractions from a<br />

45


sample (33)-(35). The selected area electron diffraction in an electron microscope is formed by<br />

setting an aperture in the imaging plane <strong>of</strong> the objective lens. Only rays passing through this<br />

aperture contribute to the diffraction pattern at the far field. For a perfect lens without<br />

aberrations, the diffracted rays come from an area that is defined by the back-projected image<br />

<strong>of</strong> the selected aperture. The aperture image is typically a factor <strong>of</strong> 20 times smaller because<br />

<strong>of</strong> the objective lens magnification. The smallest area that can be selected in SAED is limited<br />

by the objective aberration (36).<br />

2.4.2.- Scanning electron microscopy (SEM)<br />

Scanning electron microscopy (SEM) is another valuable electron microscopy technique for the<br />

examination <strong>and</strong> analysis <strong>of</strong> the microstructural characteristics <strong>of</strong> solid objects, with a<br />

resolution <strong>of</strong> ca. 5 nm. SEM can also be used to obtain compositional information <strong>of</strong><br />

nanomaterials by coupling an electron diffraction X-ray scattering detector. SEM enables the<br />

observation <strong>of</strong> heterogeneous organic <strong>and</strong> inorganic materials on the mesoscopic scale (37). In<br />

a typical SEM experiment, an electron beam is focused to obtain a very fine spot size that is<br />

rastered over the surface, <strong>and</strong> an appropriate detector collects the electrons emitted from<br />

each point. In this way, an image having a great field depth <strong>and</strong> a remarkable three-<br />

dimensional appearance is built up line by line. The specimen is usually coated with a<br />

conducting film prior to examination to enhance electron conductivity <strong>and</strong>, thus, improving<br />

image contrast.<br />

A scanning electron microscope (Figure 2.11b) consists <strong>of</strong> an electron gun at the top <strong>of</strong> the<br />

column, which creates a divergent electron beam. In the column, which is under vacuum, a<br />

series <strong>of</strong> magnetic apertures focuses the electron beam, <strong>and</strong> an electrostatic field drives the<br />

electrons through a small spot called crossover <strong>and</strong> accelerates them through the column until<br />

the sample chamber, where the electron beam interacts with the sample. The signals resulting<br />

from the beam-sample interaction are monitored. Finally, SEM constructs a virtual image from<br />

the signal emitted from the sample by scanning the electron beam line by line through a<br />

rectangular (raster) pattern on the sample surface. The scan pattern defines the area<br />

represented in the image. At any time, the beam illuminates only a single point in the pattern.<br />

As the beam moves, the signals it generates vary in strength, reflecting<br />

structural/morphological differences in the sample.<br />

46


2.5.- Atomic force microscopy (AFM)<br />

An atomic force microscope (AFM) is part <strong>of</strong> a large family <strong>of</strong> instruments termed as scanning<br />

probe microscopes (SPM). The common factor in all SPM techniques is the use <strong>of</strong> a very sharp<br />

tip probe, which is scanned across a surface <strong>of</strong> interest. The interactions between the probe<br />

<strong>and</strong> the surface are able to produce a high resolution image <strong>of</strong> the sample (potentially up to<br />

the sub-nanometre scale) depending on the technique <strong>and</strong> sharpness <strong>of</strong> the probe tip. For<br />

AFM, the probe usually interacts directly with the surface probing the repulsive <strong>and</strong> attractive<br />

forces which exist between the probe <strong>and</strong> the sample surface to produce a high resolution<br />

three-dimensional topographic image <strong>of</strong> the latter. The great versatility <strong>of</strong> AFM makes possible<br />

measurements in air or fluid environments rather than in high vacuum, which allows the<br />

imaging <strong>of</strong> polymeric <strong>and</strong> biological samples in their native states. In addition, it is highly<br />

adaptable, with tip probes being able to be chemically functionalised to allow quantitative<br />

measurement <strong>of</strong> interactions between many different types <strong>of</strong> materials (38).<br />

Figure 2.12. Typical AFM setup. A tip probe is mounted at the apex <strong>of</strong> a flexible cantilever,<br />

made <strong>of</strong> Si or Si3N4. The cantilever itself or the sample surface is mounted on a piezo-crystal,<br />

which allows the position <strong>of</strong> the probe to be shifted respect to the surface. The deflection <strong>of</strong><br />

the cantilever is monitored by changes in the path <strong>of</strong> a laser light beam deflected from the<br />

upper side end <strong>of</strong> the cantilever recorded by a photodetector (38).<br />

An AFM instrument consists <strong>of</strong> a sharp tip probe mounted near the end <strong>of</strong> a cantilever arm,<br />

which is able to make a full scanning across the entire sample surface. By means <strong>of</strong> monitoring<br />

the arm’s deflection originated by the topographic features present on the sample surface, a<br />

three dimensional picture can be built up at high resolution. In Figure 2.12 the basic setup <strong>of</strong> a<br />

47


typical AFM is shown. Cantilevers commonly are either V-shaped or a rectangular “diving<br />

board” shaped. The cantilever posses a sharp tip at its free end, which acts as the interaction<br />

probe. This probe has most commonly the form <strong>of</strong> a square-based pyramid or cylindrical cone.<br />

The probe is brought into <strong>and</strong> out <strong>of</strong> contact with the sample surface by the use <strong>of</strong> a piezo-<br />

crystal where the cantilever or the surface itself is mounted, depending on the particular<br />

equipment used. The movement in this direction is conventionally referred to as the Z-axis. A<br />

laser light beam is reflected from the reverse side <strong>of</strong> the cantilever onto a position-sensitive<br />

photodetector. The configuration <strong>of</strong> the photodetector consisted <strong>of</strong> a quadrant photodiode<br />

divide into four parts with a horizontal <strong>and</strong> vertical dividing line. If each section <strong>of</strong> the detector<br />

is labelled from A to D, as shown in Figure 2.12, then the deflection signal is calculated by the<br />

signal difference detected in each part <strong>of</strong> the detector (38).<br />

2.6.- Fluorescence spectroscopy<br />

Once a molecule is excited by the absorption <strong>of</strong> a photon, this can return its ground state<br />

through either fluorescence emission, internal energy conversion (i.e. direct return to the<br />

ground state without fluorescence emission for example, by heat radiation), intersystem<br />

crossing (possibly followed by phosphorescence emission), intramolecular charge transfer<br />

<strong>and</strong>/or conformational change. All these processes can be visualized in the Perrin-Jablonski<br />

diagram (Figure 2.13). The singlet electronic states are denoted as S0 (fundamental electronic<br />

state), S1, S2, ... <strong>and</strong> the triplet states as, T1, T2, ..., with different vibrational levels associated<br />

with each electronic state. It is important to note that energy absorption is very fast (≈10 -15 s)<br />

with respect to all other processes (so that there is no concomitant displacement <strong>of</strong> the nuclei<br />

according to the Franck-Codon principle) (39)(40). The vertical arrows corresponding to the<br />

energy absorption process start from the 0 vibrational energy level, S0, since most <strong>of</strong> molecules<br />

are in this state at room temperature. Absorption <strong>of</strong> a photon, hence, can bring a molecule to<br />

one <strong>of</strong> the vibrational levels <strong>of</strong> S1, S2, ...<br />

Emission <strong>of</strong> photons accompanying the S1 S0 relaxation is called fluorescence. It should be<br />

emphasized that, apart from a few exception, fluorescence emission occurs from S1 <strong>and</strong>,<br />

therefore, its characteristics (except polarization) do not depend on the excitation wavelength<br />

(<strong>of</strong> course only one species exists in the ground state). The transition between the ground<br />

state <strong>and</strong> the excited stated (0-transition) is usually the same for absorption <strong>and</strong> fluorescence.<br />

However, the fluorescence spectrum is located at higher wavelengths than the absorption one<br />

48


ecause <strong>of</strong> the energy loss in the excited state due to vibrational relaxation (Figure 2.14).<br />

According to the Stokes rule the fluorescence emission wavelength should be always higher<br />

than that <strong>of</strong> absorption. However, in most cases the absorption spectrum partly overlaps the<br />

fluorescence spectrum, i.e. a fraction <strong>of</strong> light is emitted at shorter wavelengths than the<br />

absorbed light. Such an observation seems to be, at first, in contradiction with the energy<br />

conservation principle. However, such energy defect is compensated by the fact that a small<br />

fraction <strong>of</strong> molecules is in a higher vibrational level in the ground state as well as in the excited<br />

state at room temperature (39).<br />

Figure 2.13. Perrin-Jablonski diagram <strong>and</strong> illustration <strong>of</strong> the relative positions <strong>of</strong> absorption,<br />

fluorescence <strong>and</strong> phosphorescence processes (39).<br />

In general, differences between the vibrational levels are similar in the ground <strong>and</strong> excited<br />

states, so that the fluorescence spectrum <strong>of</strong>ten resembles the first absorption b<strong>and</strong>. The gap,<br />

expressed in wavenumber, between the maximum <strong>of</strong> the first absorption b<strong>and</strong> <strong>and</strong> the<br />

maximum <strong>of</strong> fluorescence is called the Stokes shift.<br />

It should be noted that photon emission is as fast as photon absorption. However, excited<br />

molecules remain in the S1 state for a certain time (a few tens <strong>of</strong> picoseconds to a few<br />

hundreds <strong>of</strong> nanoseconds, depending on the type <strong>of</strong> molecule <strong>and</strong> its surrounding medium)<br />

before emitting a photon or undergoing other relaxation processes. Thus, after excitation <strong>of</strong> a<br />

49


population <strong>of</strong> molecules by a very short light pulse, the fluorescence intensity decreases<br />

exponentially with a characteristic time, reflecting the average lifetime <strong>of</strong> the molecules in the<br />

S1 excited state.<br />

Figure 2.14. Illustration <strong>of</strong> the vibrational b<strong>and</strong>s in the absorption <strong>and</strong> fluorescence spectra.<br />

As a consequence <strong>of</strong> the strong influence <strong>of</strong> the local environment or surrounding medium on<br />

fluorescence emission, fluorescent molecules are currently used for physicochemical,<br />

biochemical <strong>and</strong> biological investigation. For example, fluorescence spectroscopy can be<br />

employed as a highly sensitive method for protein analysis. Intrinsic protein fluorescence<br />

deriving from the naturally fluorescent amino acid tryptophan (<strong>and</strong> to a lesser extent from<br />

tyrosine <strong>and</strong> phenylalanine) can provide information on the protein conformational structure<br />

<strong>and</strong> it changes. On the other h<strong>and</strong>, the use <strong>of</strong> extrinsic fluorescent dyes <strong>of</strong>fer additional<br />

possibilities for protein characterization, i.e. to monitor their refolding <strong>and</strong> unfolding<br />

processes, to detect molten globule intermediates, or to characterize protein aggregation <strong>and</strong><br />

fibrillation amongst others. For example, the dye 1-anilinonaphthalene-8-sulfonate (ANS) is<br />

hardly fluorescence in aqueous environments but become highly fluorescent in apolar-organic<br />

solvents or upon adsorption to solid phases. Other important dyes for protein characterization<br />

are, for example, Nile Red (used for monitoring protein conformational changes), 9-<br />

(dicyanovinyl)-julolidine (DCVJ, used to probe tubulin structure <strong>and</strong> the formation <strong>of</strong><br />

hydrophobic microdomains along protein aggregation), or Congo Red <strong>and</strong> Thi<strong>of</strong>lavin T (used to<br />

characterize the extent <strong>of</strong> -sheet formation along protein fibrillation) (40)(41).<br />

50


Figure 2.15. Schematic representation <strong>of</strong> a spectr<strong>of</strong>luorometer (40).<br />

Figure 2.15 shows the components <strong>of</strong> a conventional spectr<strong>of</strong>luorometer. The light source is<br />

commonly a high-pressure xenon arc lamp, which <strong>of</strong>fers the advantage <strong>of</strong> a continuous<br />

emission from 250 nm to the infrared. A monochromator is used to select the excitation<br />

wavelength. Fluorescence is collected at right angles with respect to the incident beam <strong>and</strong><br />

detected through a monochromator by a photomultiplier. Automatic scanning <strong>of</strong> wavelengths<br />

is achieved by motorized monochromators, which are controlled by the electronic devices <strong>and</strong><br />

the computer, in which the data are stored (39)(40). The optical module contains various parts:<br />

a sample holder, shutters, polarizers if necessary, <strong>and</strong> a beam splitter consisting <strong>of</strong> a quartz<br />

plate reflecting a few per cent <strong>of</strong> the exciting light towards a quantum counter or a<br />

photodiode. A quantum counter usually consists <strong>of</strong> a triangular cuvette, containing a<br />

concentrated solution <strong>of</strong> a dye whose fluorescence quantum yield is independent <strong>of</strong> the<br />

excitation wavelength.<br />

51


2.7.- Circular dichroism (CD)<br />

Circular dichroism (CD) spectroscopy is commonly used for the study <strong>of</strong> macromolecules<br />

(synthetic <strong>and</strong> natural). CD technique is based on the rotation <strong>and</strong> absorption <strong>of</strong> left <strong>and</strong> right<br />

circularly polarized light passing through a sample so < CD signal will be generated when the<br />

circularly polarized light pass through a chiral sample. In this regard, a plane-polarised light<br />

beam can be considered to consist <strong>of</strong> two equal <strong>and</strong> opposing left <strong>and</strong> right circularly polarized<br />

light beams. In Figure 2.16, the electric vector <strong>of</strong> these beams is viewed looking down the<br />

direction <strong>of</strong> the light source. When the magnitudes <strong>of</strong> these vectors are equal, their resultant<br />

will trace the linear path <strong>of</strong> the electric vector amplitude <strong>of</strong> a plane-polarised light beam. This<br />

means that the magnitudes <strong>of</strong> the vector will be no longer the same. This difference involves<br />

that the resultant vector will be that <strong>of</strong> an elliptically polarized light beam passing through the<br />

sample (40)(42):<br />

Figure 2.16. Illustration <strong>of</strong> how circularly polarized light interacts with achiral <strong>and</strong> chiral<br />

molecules. Plane polarized light can be considered as composed <strong>of</strong> the sum <strong>of</strong> two equal <strong>and</strong><br />

opposing left <strong>and</strong> right circularly beams. These views represent the electric vector component<br />

<strong>of</strong> the beams looking down the direction <strong>of</strong> propagation <strong>of</strong> the light source. a) The resultant<br />

transmitted beam is still present when the plane polarized light pass through an achiral<br />

sample; b) circularly polarized light is present when it passes through a chiral sample; <strong>and</strong> c)<br />

pictorial representation defining ellipticity (42).<br />

In many biological systems, the presence <strong>of</strong> different asymmetric chemical groups around a<br />

carbon atom originates chirality, as in the case <strong>of</strong> amino acids forming the polypeptide<br />

52


ackbone <strong>of</strong> proteins. Since proteins, then, are built by units that are chiral, they will interact<br />

with left <strong>and</strong> right circularly polarized light to different extents giving rise to circular dichroism<br />

(40)(42). Hence, circular dichroism can be defined as the difference between the absorption <strong>of</strong><br />

left <strong>and</strong> right circularly polarized light by a chiral molecule:<br />

where AL <strong>and</strong> AR are the absorption <strong>of</strong> the left <strong>and</strong> right circularly polarized light, respectively,<br />

by the chiral molecule. For achiral molecules , <strong>and</strong> no CD signal is present. For chiral<br />

molecules, at some wavelength, <strong>and</strong> the resulting will be non-zero; it can have<br />

either a positive or negative value depending on the relative intensities <strong>of</strong> the left <strong>and</strong> right-<br />

h<strong>and</strong>ed light absorbance. Usually, in the analysis <strong>of</strong> protein structures CD instruments scale the<br />

chirality as ellipticity, , expressed in millidegrees. Therefore, a CD signal can be viewed either<br />

as an absorption difference or as the produced ellipticity, although both are related each<br />

other. However, ellipticity is not a useful quantity to compare different samples as its<br />

magnitude differs depending on the amount <strong>of</strong> protein present in the sample. For CD spectral<br />

data, the most useful units are delta epsilon the molar circular dichroism, Δ, <strong>and</strong> the mean<br />

residue ellipticity []MRE (MRE).<br />

Figure 2.17. Schematic representation <strong>of</strong> a typical circular dichroism spectrometer.<br />

To obtain a circular dichroism spectrum (Figure 2.17), a periodic variation in the polarization <strong>of</strong><br />

the light beam is induced by the polarization modulator through all ellipticities. This polarized<br />

light passes through the sample to a photomultiplier detector. If the sample is not optically<br />

2.32<br />

53


active, the light beam does not vary through this cycle. With the introduction <strong>of</strong> an optically<br />

active sample, a preferential absorption is seen during one <strong>of</strong> the polarization periods, <strong>and</strong> the<br />

intensity <strong>of</strong> the transmitted light now varies during the modulation cycle. The variation is<br />

directly related to the circular dichroism <strong>of</strong> the sample at that specific wavelength. Successive<br />

detection is performed at various wavelengths leading to the generation <strong>of</strong> the full CD<br />

spectrum.<br />

2.8.- Infrared spectroscopy<br />

Molecules consist <strong>of</strong> atoms which have a certain mass <strong>and</strong> are connected by elastic bonds. As a<br />

result, they can perform periodic motions, i.e. they have vibrational degrees <strong>of</strong> freedom. All<br />

motions <strong>of</strong> the atoms in a molecule relative one each other are a superposition <strong>of</strong> the so-called<br />

normal frequencies. These are described by a normal coordinate. For polyatomic molecules,<br />

the number <strong>of</strong> characteristic frequencies is for non-linear molecules <strong>and</strong> for<br />

linear molecules, where N is the number <strong>of</strong> atoms in the molecule, which define their<br />

vibrational spectra. These spectra depend on atoms mass, their geometrical arrangement, <strong>and</strong><br />

the strength <strong>of</strong> their chemical bonds. Molecular aggregates such as crystals or complexes<br />

behave like “super-molecules” in which the vibrations <strong>of</strong> the individual components are<br />

coupled. As a first approximation the normal vibrations are considered not coupled, i.e. they<br />

do not interact. However, the bonds elasticity does not strictly follow Hooke’s law. Therefore,<br />

overtone <strong>and</strong> combinations <strong>of</strong> normal vibrations may appear (40)(43).<br />

Infrared (<strong>and</strong> RAMAN) spectrometers are the most important tools for observing vibrational<br />

spectra. Depending on the nature <strong>of</strong> the vibration, which is determined by the symmetry <strong>of</strong><br />

the molecule, vibration may be active or forbidden in the infrared (or RAMAN) spectrum.<br />

From basic theory, it is easy to underst<strong>and</strong> that a molecule can take up an amount <strong>of</strong> energy,<br />

, to reach a vibrationally excited state only when the frequency <strong>of</strong> the incident radiation on<br />

a molecule is the same as the vibrational frequency <strong>of</strong> the latter. This phenomenon involves<br />

either a change <strong>of</strong> the molecule’s dipole moment or a change in its polarizability. If the former<br />

phenomenon takes place, we will observe an IR b<strong>and</strong>; if the second one occurs, we will<br />

observe a RAMAN b<strong>and</strong>. The absorption wavelength range <strong>of</strong> IR spectroscopy is from 2.5 to<br />

1000 μm, which corresponds to the energy with = 4000-10 cm -1 , with the light<br />

velocity.<br />

54


According to classical mechanics, the vibrating atoms groups in a molecule can be modelled as<br />

balls joined by springs (the molecular bonds). To take the simplest example, let us consider<br />

two balls joined by one spring representing a diatomic molecule. In this approximation, two<br />

atoms <strong>and</strong> the connecting bond are treated as a simple harmonic oscillator composed <strong>of</strong> 2<br />

masses (atoms or balls). These two balls can vibrate with a characteristic frequency that<br />

depends on the mass <strong>of</strong> the balls <strong>and</strong> on the spring force (spring force can be approximated by<br />

Hooke’s law). Matematically, the equation<br />

, describes the characteristic<br />

oscillation frequency <strong>and</strong> how it depends on mass <strong>and</strong> spring constant force . According<br />

to this equation, stiffer spring constant (stronger bonds) result in a higher frequency, whereas<br />

heavier atoms result in lower frequencies. This simple system possess only a single<br />

characteristic frequency. The balls can only vibrate with this characteristic frequency or an<br />

integer multiple <strong>of</strong> this frequency. A more complex molecule can still be modelled quite<br />

accuarately as balls <strong>of</strong> different massess joined by springs <strong>of</strong> different tensions (representing<br />

different atoms with different bonds). Figure 2.18 shows a diagram representing the vibration<br />

<strong>of</strong> a diatomic molecule or one vibrational degree <strong>of</strong> freedom <strong>of</strong> a polyatomic molecule, which<br />

is excited by a transition from energy level to (40)(43)-(45).<br />

Figure 2.18. Observation <strong>of</strong> the excitation <strong>of</strong> a vibrational state in the electronic ground state<br />

by infrared absorption.<br />

In the classical harmonic oscillator,<br />

, where x is the displacement <strong>of</strong> the spring.<br />

Thus, the energy or frequency is dependent on how far one stretches or compresses the<br />

spring. If this simple model were true, a molecule could absorb energy <strong>of</strong> any wavelength.<br />

However, vibrational motion is quantized <strong>and</strong>, therefore, the molecules can only exist in finite<br />

energy states. In the case <strong>of</strong> harmonic potentials, these states are equidistant, i.e.<br />

where the energy levels are numbered as ni. At ni = 0, the<br />

potential energy has its lowest value, which is not the potential minimum energy (43)(44).<br />

55


According to the selection rules, only transitions to the next energy level are allowed;<br />

therefore, molecules will absorb an amount <strong>of</strong> energy equal to . This rule is not inflexible<br />

<strong>and</strong>, occasionally, transitions <strong>of</strong> , , or higher are observed. These correspond to b<strong>and</strong>s<br />

called overtones in an IR spectrum. They are <strong>of</strong> lower intensity than the fundamental vibration<br />

b<strong>and</strong>s.<br />

As mentioned above, infrared spectra cover the near infrared(0.75-3 m), middle infrared (3-<br />

30 m) <strong>and</strong> far infrared (30-300 m) range. In this region, most molecules show absorption or<br />

emission b<strong>and</strong>s arising from symmetry-allowed interaction with the radiation. The energy <strong>of</strong><br />

the absorbed <strong>and</strong> emitted light is equivalent to the energy difference between the lowest<br />

vibrational states <strong>of</strong> the electronic ground state <strong>of</strong> a molecule.<br />

Three types <strong>of</strong> instruments for IR absortion measurements are commonly available: dispersive<br />

spectrophotometers with a grid monochromator, Fourier thansform spectrometers employing<br />

an interferometer, <strong>and</strong> non-dispersive photometers using a filter or absorbing gas (typically<br />

used for analysis <strong>of</strong> atmospheric gases at specific wavelengths).<br />

Until 1980s, the most widely used instruments for IR measurements were dispersive<br />

spectrometers. Nowadays, this type <strong>of</strong> instruments has been displaced by Fourier transform<br />

(FT) spectrometers because <strong>of</strong> their speed, reability, enhanced signal-to-noise ratio. Infrared<br />

spectrometers (Figure 2.19) usually combine a radiation source, collimator, fixed mirror,<br />

moving mirror, a sample holder, a device for spectral dispersion or selection <strong>of</strong> radiation<br />

(beam splitter), <strong>and</strong> a radiation detector, all connected to appropiate recording <strong>and</strong> evaluation<br />

facilities. An ideal spectrometer records completely resolved spectra with a maximum signal-<br />

to-noise ratio. It requires a minimum amount <strong>of</strong> sample which is measured nondestructively in<br />

a minimum time, <strong>and</strong> it provides facilities for storing <strong>and</strong> evaluating the spectra. FTIR<br />

spectrometers can be single-beam or double-beam instruments. A typical procedure for<br />

determining transmittance or absorbance with this type <strong>of</strong> instrument is to first obtain a<br />

reference interferogram by scanning a reference (usually air) twenty or thirty times coadding<br />

the data, <strong>and</strong> storing the results in the computer. A sample is inserted in the radiation path<br />

<strong>and</strong> the process repeated. The ratio <strong>of</strong> the sample to reference spectral data is, then,<br />

computed to obtain the transmittance at various frequencies.<br />

Infrared spectroscopy (IR) has long been used to determine molecular structures <strong>and</strong> to<br />

identify unknown compounds. IR data have been used to obtain information about the<br />

56


chemical composition, configuration, <strong>and</strong> crystallinity <strong>of</strong> polymeric <strong>and</strong> biological materials<br />

(43). In particular, FTIR spectroscopy is a versatile spectroscopic technique for analyzing<br />

protein structure in solution, in aggregates <strong>and</strong> inclusion bodies. Most IR studies focus on the<br />

characterization <strong>of</strong> protein secondary structures by using absorption b<strong>and</strong>s reflecting the<br />

motions <strong>of</strong> peptide groups. Infrared spectra <strong>of</strong> proteins may be obtained either in the solid<br />

state or in solution. The vibrational spectra <strong>of</strong> proteins exhibit several relatively intense peaks<br />

which shift in wavenumber <strong>and</strong> intensity from one sample to another. There are several<br />

prominent b<strong>and</strong>s for the study <strong>of</strong> polypeptides <strong>and</strong> proteins; however in this study, we will<br />

focused partularly on the amide b<strong>and</strong>s, which are <strong>of</strong> particular interest for the study <strong>of</strong> protein<br />

conformation: the amide I, between 1680 <strong>and</strong> 1600 cm -1 , is mainly associated with the<br />

stretching vibration <strong>of</strong> the C=O bond; amide II, found between 1580 <strong>and</strong> 1480 cm -1 , <strong>and</strong> amide<br />

III b<strong>and</strong>, observed between 1300 <strong>and</strong> 1200 cm -1 , are both associated with coupled C-N<br />

stretching <strong>and</strong> N-H bending vibrations <strong>of</strong> the peptide groups, respectively. These b<strong>and</strong>s are<br />

conformationally sensitive, <strong>and</strong> their position may be attributed to a dominat protein<br />

conformation, that is, α-helix, parallel <strong>and</strong> antiparallel β-sheets, turns or r<strong>and</strong>om coil (38)(42).<br />

In particular, the following vibrational frequencies had been assigned to: 1654 cm -1 for α-helix;<br />

(1624, 1631, 1637, <strong>and</strong> 1675 cm -1 for β-sheets; 1663, 1670, 1683, 1688, <strong>and</strong> 1694 cm -1 for<br />

turns; <strong>and</strong> 1645 cm -1 for other structures (43)(46)-(48). The relative amounts <strong>of</strong> the four<br />

secondary structures are determined by the relative areas under the b<strong>and</strong> peaks.<br />

Figure 2.19. Scheme <strong>of</strong> a FTIR instrument.<br />

57


2.9.- Rheology<br />

Rheology involves the study <strong>of</strong> the deformation <strong>and</strong> flow <strong>of</strong> matter due to compressive<br />

stresses acting on a material. Particularly, it refers to the behaviour <strong>of</strong> materials when a<br />

mechanical force is applied on. Rheology includes three main concepts such as force,<br />

deformation <strong>and</strong> time. Irreversible flows, reversible elastic deformations or their combination<br />

(viscoelasticity) can, therefore, model <strong>and</strong> describe a rheological phenomenon under certain<br />

assumptions. The type <strong>of</strong> deformation depends on the state <strong>of</strong> matter; for example, gases <strong>and</strong><br />

liquids will flow when a force is applied, whilst solids will deform by a fixed amount <strong>and</strong>, then,<br />

back to their original shape when the force is removed. In the case <strong>of</strong> polymers, the rheological<br />

<strong>and</strong> mechanical properties affect the polymers molecular properties, such as their molecular<br />

weight, molecular weight distribution, conformation, architecture <strong>and</strong> crystallinity (49)(50).<br />

Figure2.20 Schematic diagram <strong>of</strong> basic tool geometries for a rotational rheometer: a)<br />

concentric cylinder, b) cone <strong>and</strong> plate, c) parallel plate (50).<br />

A typical rheometer measures the velocity <strong>of</strong> displacement <strong>of</strong> the moving surface <strong>and</strong> the<br />

force exerted on one <strong>of</strong> the surfaces. Most rheometers are based on rotary motion <strong>and</strong> use<br />

one <strong>of</strong> the three following geometries (Figure 2.20): concentric cylinder, cone <strong>and</strong> plate, <strong>and</strong><br />

parallel disk. In most cases the same rotary instrument can use all three <strong>of</strong> these flow<br />

geometries. To generate the needed motion, they typically use actuators like a hydraulic<br />

piston or ball screws found in st<strong>and</strong>ard tensile testing machines for solids. Solenoids or other<br />

electromechanical actuator are <strong>of</strong>ten used for small amplitudes <strong>and</strong> low forces. There are two<br />

basic designs <strong>of</strong> rheometers: controlled stress ones, where the stress is applied electrically via<br />

a motor measuring the strain; <strong>and</strong> controlled strain instruments, in which a strain is imposed<br />

<strong>and</strong> the stress is computed from the deformation <strong>of</strong> a calibrated spring system (50).<br />

58


2.9.1.- Viscoelasticity<br />

Many materials can be classified as solid or fluids, displaying elastic <strong>and</strong> viscous behaviour<br />

respectively. Viscoelastic materials such as polymers combine the characteristics <strong>of</strong> both<br />

elastic <strong>and</strong> viscous materials depending on the experimental time scale. Application <strong>of</strong> a stress<br />

<strong>of</strong> relatively long duration may cause some flow <strong>and</strong> irrecoverable deformation, while a rapid<br />

shearing would induce an elastic response in some polymeric fluids. Then, a classification <strong>of</strong><br />

these materials should include a consideration <strong>of</strong> the timescale <strong>of</strong> the measurement relative to<br />

the characteristic time <strong>of</strong> the material. This ratio is given by the Deborah number (De): when De<br />

is 1 the material will display both viscous <strong>and</strong> elastic behaviour <strong>and</strong> is described as viscoelastic,<br />

whilst, for De » 1 <strong>and</strong> De « 1 the material has solid-like <strong>and</strong> viscous-like behaviour, respectively<br />

(49)(50).<br />

Pure elastic solid behaviour may be exemplified by a Hookean spring, <strong>and</strong> pure viscous flow<br />

can be exemplified by the behaviour <strong>of</strong> a dashpot, which is essentially a piston moving in a<br />

cylinder <strong>of</strong> Newtonian fluid. Nevertheless, the viscoelasticity cannot be described accurately by<br />

neither spring nor dashpot alone, but a combination <strong>of</strong> both. In this way, several mechanical<br />

models have been developed to describe the viscoelastic response <strong>of</strong> a given material. All <strong>of</strong><br />

these mechanical models are, then, supported on the combination <strong>of</strong> these two simple<br />

elements (spring <strong>and</strong> dashpot), which represented the elastic behaviour <strong>of</strong> a solid body, <strong>and</strong><br />

the viscous flow for a liquid, respectively. Among all models, Maxwell’s <strong>and</strong> Kelvin-Voigt’s<br />

models are the most frequently used (Figure 2.21) (50).<br />

a) b)<br />

<br />

E<br />

<br />

Figure 2.21. a) Maxwell’s <strong>and</strong> b) Voigt’s models.<br />

The use <strong>of</strong> mechanical models such as the spring <strong>and</strong> dashpot as analogues <strong>of</strong> the behaviour <strong>of</strong><br />

real materials enables us to describe very complex experimental behaviours using a simple<br />

combination <strong>of</strong> models. The uses <strong>of</strong> models to represent a wide range <strong>of</strong> deformation are not<br />

restricted to the application <strong>of</strong> shear stress; however, in this section, we will focus on the shear<br />

properties <strong>of</strong> the material (1).<br />

E<br />

<br />

59


To overcome the poor description <strong>of</strong> real polymeric materials by either the spring or the<br />

dashpot, Maxwell suggested a simple series combination <strong>of</strong> both elements (Figure 2.21a).<br />

Because the material possesses the ability to flow, some inertial relaxation will occur <strong>and</strong> less<br />

force will be required upon time to sustain the deformation. The goal in the Maxwell’s model<br />

is to calculate how the stress varies with time, or expressing the stress in terms <strong>of</strong> the constant<br />

strain to describe the time-dependent modulus. Let us consider in the Maxwell’s model, in<br />

which one spring is attached to one dashpot, as shown in Figure 2.21a. When a force is acting<br />

on the spring downward at (in one dimensional flow), the stress-strain relation for the<br />

spring (Hookean material) may be described by (50):<br />

where is the applied stress, is the strain, <strong>and</strong> G the elastic modulus. Conversely, the stress<br />

response <strong>of</strong> the dashpot (a purely viscous Newtonian fluid) to an applied deformation rate may<br />

be described as:<br />

where<br />

2.33<br />

2.34<br />

is the strain rate <strong>and</strong> , defines the viscous response <strong>of</strong> the dashpot with<br />

Newtonian behaviour. In other words, the spring purely exhibits an elastic effect (as a Hookean<br />

solid) <strong>and</strong> the dashpot exhibits purely a viscous effect (as a viscous fluid).<br />

In the Maxwell’s element, both the spring <strong>and</strong> the dashpot support the same stress <strong>and</strong>,<br />

therefore:<br />

where <strong>and</strong> are the stresses on the spring <strong>and</strong> dashpot, respectively. However, the<br />

overall strain <strong>and</strong> the strain rates are the sum <strong>of</strong> the elemental strain <strong>and</strong> strain rates,<br />

respectively, that is:<br />

or<br />

2.35<br />

2.36<br />

60


where is the total strain rate, while<br />

<strong>and</strong><br />

2.37<br />

are the strain rates <strong>of</strong> the spring <strong>and</strong><br />

dashpot, respectively. Therefore, the total strain <strong>of</strong> the spring <strong>and</strong> dashpot at any time is the<br />

sum <strong>of</strong> that <strong>of</strong> the spring <strong>and</strong> <strong>of</strong> the dashpot. Then, for a Maxwell’s model the strain rates can<br />

be written as:<br />

By rheometry we can be measure the response <strong>of</strong> a material to an oscillating stress or strain,<br />

so it is considered as mechanical spectroscopy. When a sample is constrained in, for example,<br />

a cone <strong>and</strong> plate assembly, an oscillating strain at a given frequency can be applied to the<br />

sample. After an initial start-up period due to a transient sample state a stress develops in<br />

direct response to the applied strain. If the strain has an oscillating value with time the stress<br />

must also be oscillating. We can represent these two wave forms as in Figure 2.22. The elastic<br />

<strong>and</strong> viscous effects are out <strong>of</strong> phase one each other by some angle (1).<br />

Figure 2.22. An oscillating strain <strong>and</strong> the stress response for a viscoelastic material.<br />

All the information about the response <strong>of</strong> the sample at this frequency is contained within<br />

these wave forms. However, this information is not in a particularly tractable form. What we<br />

would really prefer is to have a few representative terms such as the relaxation time <strong>and</strong><br />

2.38<br />

61


elasticity or viscosity <strong>of</strong> the sample in order to characterize its material properties. In order to<br />

obtain this information some mathematical manipulation is required. Two key features can be<br />

utilized. The first feature is the maximum stress, , divided by the maximum strain, , which is<br />

constant for a given frequency . This ratio is called the complex modulus :<br />

is the radial frequency, which is , where is the applied frequency measured in Hz. The<br />

other feature constant with time at any given frequency is , the phase difference (expressed<br />

in radians). These two values, <strong>and</strong> , are characteristic <strong>of</strong> the material. It is straightforward<br />

to visualise the situation where an elastic solid is placed in a cone <strong>and</strong> plate geometry. When a<br />

tangential displacement is applied to the lower plate a strain in the sample is produced. That<br />

displacement is transmitted directly through the sample. The upper cone will react<br />

proportionally to the applied strain to give a stress response. An oscillating strain will give an<br />

oscillating stress response that is in phase with the strain, so will be zero. However, if we<br />

have a Newtonian liquid, the peak stress will be out <strong>of</strong> phase by<br />

2.39<br />

rad as the peak stress is<br />

proportional to the strain rate. So, it follows that if we have a viscoelastic material some<br />

energy is stored <strong>and</strong> some dissipated, <strong>and</strong> the stored contribution will be in phase whilst the<br />

dissipated or loss contribution will be out <strong>of</strong> phase respect to the applied strain.<br />

In order to describe the material properties as a function <strong>of</strong> frequency we need to use the<br />

constitutive equation 2.30. This equation describes the relation between the stress <strong>and</strong> the<br />

strain. However, it is most convenient to express the applied sinusoidal wave in the<br />

exponential form <strong>of</strong> the complex number notation:<br />

Now, the stress response lags by the phase angle :<br />

So, substituting the complex stress <strong>and</strong> strain into the constitutive equation for a Maxwell fluid<br />

the resulting relation is given by:<br />

2.40<br />

2.41<br />

62


Using equations 2.40 <strong>and</strong> 2.41 in equation 2.42 <strong>and</strong> rearranging we have:<br />

Thus, arrangement <strong>of</strong> this expression gives the complex modulus <strong>and</strong> frequency:<br />

or:<br />

where<br />

2.42<br />

2.43<br />

2.44<br />

2.45<br />

is the characteristic relaxation time. This expression describes the variation <strong>of</strong><br />

the complex modulus with frequency for the Maxwell model. It is normal to separate the real<br />

<strong>and</strong> imaginary components <strong>of</strong> this expression. This is achieved by multiplying by ( ) to<br />

give:<br />

where is an in phase elastic modulus with energy storage in the periodic deformation,<br />

<strong>and</strong> is called the dynamic storage modulus. is an out <strong>of</strong> phase elastic modulus associate<br />

with the energy dissipation as heat, <strong>and</strong> is called the dynamic loss modulus. Then:<br />

These expressions describe the frequency dependence <strong>of</strong> the stress with respect to the strain.<br />

2.46<br />

2.47<br />

2.48<br />

63


It is normal to represent them as two moduli that determine the component <strong>of</strong> stress in phase<br />

with the applied strain (storage modulus) <strong>and</strong> the component out <strong>of</strong> phase by 90 0 . In an<br />

experiment, the amplitudes <strong>of</strong> the oscillation input ( ) <strong>and</strong> output ( ), <strong>and</strong> the phase angle<br />

( ) are measured. Therefore, each oscillatory shear flow measured at a given provides two<br />

independent quantities, amplitude ratio <strong>and</strong> phase angle:<br />

The storage <strong>and</strong> loss moduli are subtle descriptions <strong>of</strong> the material properties <strong>of</strong> the system.<br />

These two properties are related to the phase angle <strong>and</strong> complex modulus. These are both<br />

functions <strong>of</strong> the applied frequency <strong>and</strong> represent an alternative description <strong>of</strong> the system.<br />

The phase angle changes with the frequency, from 90 degrees at low frequency, to 0 degrees<br />

at the high frequency limit; thus, as the frequency increases the sample becomes more elastic,<br />

<strong>and</strong> the phase difference between the stress <strong>and</strong> the strain reduces.<br />

2.10.- X-ray diffraction<br />

X-ray diffraction (XRD) is a versatile, powerful <strong>and</strong> non destructive technique that reveals<br />

detailed information about the chemical crystallographic structure <strong>of</strong> natural <strong>and</strong><br />

manufactured materials. Diffraction effects are observed when an electromagnetic radiation<br />

impinges on periodic structures with geometrical variations on the length scale <strong>of</strong> the radiation<br />

wavelength. The inter-atomic distances in crystal <strong>and</strong> molecules is around to 0.15 – 0.5 nm,<br />

which correspond in the electromagnetic spectrum with the wavelength <strong>of</strong> X-rays, having<br />

photon energies between 3 <strong>and</strong> 8 keV (3).<br />

There are three different types <strong>of</strong> interactions between X-rays <strong>and</strong> matter. First, the<br />

photoionization, where electrons may be liberated from their bound atomic states during the<br />

process, <strong>and</strong> which falls into the group <strong>of</strong> inelastic processes. In addition, other inelastic<br />

2.49<br />

2.50<br />

2.51<br />

64


scattering process that incoming X-ray beams may undergo upon interaction with matter is the<br />

energy transfer to an electron but, in this case, without a release <strong>of</strong> the electron from the<br />

atom (phenomenon known as Compton scattering). Finally, X-rays may be elastically scattered<br />

by electrons, process known as Thomson scattering. In this process, the electron oscillates at<br />

the frequency <strong>of</strong> the incoming beam so the wavelength <strong>of</strong> the scattered radiation is conserved.<br />

There are many theories <strong>and</strong> equations to correlate the diffraction pattern <strong>and</strong> the material<br />

structure. The Bragg law is a simple way to describe the diffraction <strong>of</strong> X-rays by a crystal. In<br />

Figure 2.23a, the incident X-rays reach the crystal planes with an incident angle θ <strong>and</strong> are<br />

diffracted at an angle θ. The diffraction peak is observed when the Bragg condition is satisfied<br />

(3)(51):<br />

where λ is the wavelength <strong>of</strong> the incident X-rays, d is the distance between each adjacent<br />

crystal plane (d-spacing), θ is the Bragg angle at which a diffraction peak is observed, <strong>and</strong> n (1,<br />

2, ...) is an integer number, called the reflection order; this means that the Bragg condition can<br />

be satisfied by various X–ray wavelengths. The Bragg law gives a simple relationship between<br />

the diffraction pattern <strong>and</strong> the crystal structure, <strong>and</strong> many X-ray diffraction applications can be<br />

easily explained by this law.<br />

Figure 2.23. a) Incident <strong>and</strong> reflected X-rays form an angle θ which is symmetric to the crystal<br />

normal plane; b) The diffraction peak is observed at a Bragg angle θ (52).<br />

In X-ray diffraction using a single wavelength, the Bragg equation is typically expressed with n<br />

= 1 for the first order <strong>of</strong> diffraction because higher order reflections can be considered as<br />

2.52<br />

65


esulting from different lattice planes. For instance, the second-order reflection from (hkl)<br />

planes is equivalent to the first-order reflection from (2h, 2k, 2l) planes. The diffraction peak is<br />

displayed as diffracted intensities at a range <strong>of</strong> 2θ angles. For perfect crystals, the peak is a<br />

delta function, the dark straight vertical line shown in Figure 2.23b (52).<br />

Figure 2.24. Diffraction patterns from crystalline solids, liquids, amorphous solids, <strong>and</strong><br />

monoatomic gases as well as their mixtures (52).<br />

A typical diffraction peak is a broadened peak displayed by the curved line in Figure 2.23b. The<br />

peak broadening can be due to many effects, including imperfect crystal conditions, such<br />

strain, mosaic structures, <strong>and</strong> finite sizes; ambient conditions, such as atomic thermal<br />

vibrations; <strong>and</strong> instrumental conditions, such as X-ray beam size, beam divergence, beam<br />

spectrum distribution, <strong>and</strong> detector resolution. The curved line gives a peak pr<strong>of</strong>ile, which is<br />

the diffracted intensity distribution <strong>of</strong> the Bragg angle. The highest point on the curve gives the<br />

66


maximum intensity <strong>of</strong> the peak, Imax. The width <strong>of</strong> the peak is typically measured by its full<br />

width at half maximum (FWHM). The total diffracted energy can be measured by the area<br />

under the peak, which is referred to as integrated intensity. The integrated intensity is a more<br />

consistent value for measuring the diffracted intensity <strong>of</strong> a reflection since it is less affected by<br />

all peak broadening factors (51)(52).<br />

The Bragg diffraction condition is based on the existence <strong>of</strong> a long periodicity <strong>of</strong> crystalline<br />

materials. As mentioned above, X-rays diffraction can provide information on the atomic<br />

arrangement in material with long-range order, short-range order, or no order at all, like gases,<br />

liquid, <strong>and</strong> amorphous solids, <strong>and</strong> a material may have one or to be a mixture <strong>of</strong> the former<br />

atomic arrangement types. Figure 2.24 gives a schematic comparison <strong>of</strong> the diffraction<br />

patterns for crystalline solids, liquids, amorphous solids, <strong>and</strong> monatomic gases as well as their<br />

mixtures. The diffraction patterns shown in this Figure are displayed as diffracted intensity<br />

versus assuming that the diffracted intensity is a unique function <strong>of</strong> the diffraction angle.<br />

The diffraction pattern from crystals has many sharp peaks corresponding to various crystal<br />

planes based on the Bragg law. The peaks at low angles are from crystal planes <strong>of</strong> large d-<br />

spacing <strong>and</strong>, viceversa, at high angles. To satisfy the Bragg condition at all crystal planes,<br />

the crystal diffraction pattern is actually generated from polycrystalline materials or powder<br />

materials. Therefore, the diffraction pattern is also called powder diffraction pattern. A similar<br />

diffraction pattern can be collected with a single crystal if this has been rotated at various<br />

angles during data collection so that the Bragg law can be satisfied when the crystal is at the<br />

right orientation.<br />

Both amorphous solid <strong>and</strong> liquid materials do not have the long-range order <strong>of</strong> crystals, but<br />

their atomic distances have a narrow distribution because <strong>of</strong> atoms are tightly packed. In this<br />

case, the intensity <strong>of</strong> the scattered X-rays forms one or two maxima with a very broad<br />

distribution in the range. The intensity vs. distribution, thus, reflects the distribution <strong>of</strong><br />

the atomic distances. In the case <strong>of</strong> a gas there is no order at all, i.e. the atoms are distributed<br />

r<strong>and</strong>omly in space. The scattering curve is featureless except that the scattered intensity drops<br />

continuously with the increase <strong>of</strong> . On the other h<strong>and</strong>, the diffraction pattern from a<br />

material containing both amorphous <strong>and</strong> crystalline features has a broad background from the<br />

amorphous phase <strong>and</strong> sharp peaks from the crystalline phase. For example, many polymer<br />

materials have an amorphous matrix with crystallized regions. The diffraction pattern may<br />

contain air-scattering background in addition to sharp diffraction peaks. The air-scattering can<br />

be generated from the incident beam or diffraction beam. If the air-scattering is not removed<br />

67


y the diffractometer, the diffraction pattern contains a high background at low angle <strong>and</strong><br />

the background gradually decreases with increasing angle (52).<br />

On the other h<strong>and</strong>, X-ray diffraction phenomena can also be explained in reciprocal space by<br />

the reciprocal lattice <strong>and</strong> the Ewald sphere. Reciprocal lattice is a transformation <strong>of</strong> the crystal<br />

lattice in real space to reciprocal space. The shape <strong>and</strong> size <strong>of</strong> a unit cell in real space can be<br />

defined by three vectors, a, b, <strong>and</strong> c, all starting from any single lattice point. The unit cell <strong>of</strong><br />

the corresponding reciprocal lattice is, then, given by three vectors a*, b*, <strong>and</strong> c* (also<br />

referred to as reciprocal lattice axes), <strong>and</strong>:<br />

where V is the volume <strong>of</strong> the crystal unit cell in the real space <strong>and</strong><br />

Since each reciprocal lattice axis is the vector product <strong>of</strong> two lattice axes in real space, these<br />

are perpendicular to the plane defined by the two lattice axes. The original lattice axes <strong>and</strong><br />

reciprocal lattice axes maintain the following relations:<br />

<strong>and</strong><br />

Figure 2.26 illustrate the relationship between the original lattice in real space <strong>and</strong> the<br />

reciprocal lattice. The unit cell <strong>of</strong> the original lattice is drawn in dotted lines. The three<br />

reciprocal lattice axes define a unit cell <strong>of</strong> the reciprocal lattice (solid lines). The origin <strong>of</strong> the<br />

reciprocal lattice axes, denoted by O, is the origin <strong>of</strong> the reciprocal lattice. The repeat<br />

2.53<br />

2.54<br />

2.55<br />

2.56<br />

68


translation <strong>of</strong> the reciprocal lattice unit cell in three dimensions forms the complete reciprocal<br />

lattice. Except the origin, each lattice point is denoted by a set <strong>of</strong> three integers (hkl), which<br />

are the number <strong>of</strong> translations <strong>of</strong> the three reciprocal lattice axes, respectively, to reach the<br />

lattice point. In other words, the vector drawn from the origin <strong>of</strong> the lattice point (hkl) is given<br />

by:<br />

<strong>and</strong> the direction <strong>of</strong> the vector is normal to the lattice planes (hkl) in real space. The<br />

magnitude <strong>of</strong> the vector is given by the d-spacing <strong>of</strong> the (hkl) planes:<br />

Figure 2.26. Relationship between the original lattice in real space <strong>and</strong> the reciprocal lattice<br />

(49).<br />

Therefore, each point (hkl) in the reciprocal lattice represents a set <strong>of</strong> lattice planes (hkl) in the<br />

real space lattice. The position <strong>of</strong> the point in the reciprocal lattice defines the orientation <strong>and</strong><br />

d-spacing <strong>of</strong> the lattice planes in the real space lattice. The farther away a reciprocal lattice<br />

point is from the origin, the smaller is the d-spacing <strong>of</strong> the corresponding lattice planes. For<br />

example, the reciprocal lattice point (111) represent the (111) lattice planes in the real space<br />

lattice, <strong>and</strong> the lattice vector is given by <strong>and</strong><br />

.<br />

2.57<br />

2.58<br />

69


Figure 2.27. The Ewald sphere <strong>and</strong> Bragg condition in reciprocal space.<br />

On the other h<strong>and</strong>, the relationship between the Bragg condition <strong>and</strong> the reciprocal lattice can<br />

be explained visually by the Ewald sphere, also referred as reflection sphere. Ewald came up<br />

with a geometrical construction to help the visualization <strong>of</strong> which Bragg planes are in the<br />

correct orientation to diffract. In Figure 2.27, the diffracting crystal is located in the center <strong>of</strong><br />

the Ewald sphere, C. The radius <strong>of</strong> the Ewald sphere is defined as<br />

. The incident beam can<br />

be visualized as the vector from I to C, <strong>and</strong> the diffracted beam is the vector from C to P. Both<br />

the incident <strong>and</strong> diffracted beams form an angle θ from a set <strong>of</strong> crystal planes (hkl). The d-<br />

spacing <strong>of</strong> the crystal planes is dhkl In the Ewald sphere; both incident beam vector<br />

diffracted beam vector<br />

<strong>and</strong><br />

start at the poin C <strong>and</strong> end at the point O <strong>and</strong> P, respectively. The<br />

vector from O to P is the reciprocal lattice vector <strong>and</strong> is perpendicular to the crystal<br />

planes. The three vectors have the following relationship:<br />

<strong>and</strong> their magnitude can be expressed based on the Bragg law as:<br />

The point O is the origin <strong>of</strong> the reciprocal lattice, <strong>and</strong> the point P is the reciprocal point (hkl).<br />

The Bragg condition is satisfied only when the reciprocal lattice point falls on the Ewald sphere.<br />

2.59<br />

2.60<br />

70


For a single crystal, the chance to have a reciprocal lattice point on the Ewald sphere is very<br />

small if the crystal orientation is fixed. Multiplying both ends <strong>of</strong> equation 2.59 by the three<br />

lattice axes in real space, respectively, the Laue equations are obtained:<br />

2 .61<br />

The Laue equations establish that a periodic three-dimensional lattice produces diffraction<br />

maxima at specific angles depending on the incident beam direction <strong>and</strong> the wavelength. The<br />

Laue equations are suitable to describe the diffraction geometry <strong>of</strong> a single crystal. The Bragg<br />

law is more conveniently used for powder diffraction. Both the Bragg law <strong>and</strong> Laue equations<br />

define the diffraction condition in different formats.<br />

On the other h<strong>and</strong>, the distance between the origin <strong>of</strong> the reciprocal lattice O <strong>and</strong> the lattice<br />

point P is reciprocal to the d-spacing. The largest possible magnitude <strong>of</strong> the reciprocal lattice<br />

vector is given by<br />

. This means that the smallest d-spacing satisfying the Bragg condition is<br />

. In powder X-ray diffraction, the r<strong>and</strong>om orientation <strong>of</strong> all crystallites can take all possible<br />

orientations assuming an infinite number <strong>of</strong> crystallites. The trace <strong>of</strong> the reciprocal lattice<br />

points from all crystallites can be considered as a series <strong>of</strong> spherical surfaces with origin O as<br />

the center point. Therefore, the condition to satisfy the Bragg law is only if the d-spacing is<br />

greater than half <strong>of</strong> the wavelength. In other words, the Bragg condition can be satisfied if a<br />

reciprocal lattice point falls in a sphere <strong>of</strong> from the origin O. This sphere is called the<br />

limiting sphere for powder diffraction.<br />

X-ray diffraction systems have a variety <strong>of</strong> configurations <strong>and</strong> component options to fulfil<br />

requirements <strong>of</strong> different samples <strong>and</strong> applications. As shown in Figure 2.25, a typical XRD<br />

system normally consist <strong>of</strong> five basic components:<br />

The X-ray source produces X-rays with the required radiation energy, focal spot size,<br />

<strong>and</strong> intensity;<br />

The X-ray optics establish the primary X-ray beam to the required wavelength, beam<br />

focus size, beam pr<strong>of</strong>ile, <strong>and</strong> divergence;<br />

The Goniometer <strong>and</strong> sample stage establish <strong>and</strong> tune the geometric relation between<br />

71


primary beam, sample, <strong>and</strong> detector;<br />

The sample alignment <strong>and</strong> monitor assist users about sample positioning into the<br />

instrument <strong>and</strong> monitors the sample state <strong>and</strong> position;<br />

The area detector intercepts <strong>and</strong> records the scattering X-rays from a sample, <strong>and</strong> it<br />

saves <strong>and</strong> displays the diffraction pattern into a frame.<br />

Each <strong>of</strong> the former basic components may have several options suitable for various application<br />

<strong>and</strong> functions. The whole system is controlled by a computer equipped with a s<strong>of</strong>tware for<br />

instrument control, data acquisition <strong>and</strong> data analysis. In addition to the five basic<br />

components, there are some other accessories, such as a low <strong>and</strong> high-temperature stages,<br />

helium or vacuum beam path for SAXS, beam stop, <strong>and</strong> alignment <strong>and</strong> calibration fixtures.<br />

Figure 2.25. Scheme <strong>of</strong> a X-ray diffractometer.<br />

2.11.- Additional experimental techniques used in fibrillogenesis studies<br />

2.11.1.- UV-Vis absorption spectroscopy<br />

Spectra are observed as the result <strong>of</strong> the energy exchange between a substance <strong>and</strong> the<br />

electromagnetic radiation. If the energy is absorbed by a substance from the radiation field,<br />

we have absorption spectra; if the energy is added to the radiation field, emission spectra are<br />

observed. An electromagnetic radiation may be characterized by the frequency , the<br />

wavelength , or the wave number . They are related as ,<br />

velocity <strong>of</strong> light (39)(40)(53).<br />

; where c is the<br />

The three most important types <strong>of</strong> optical spectroscopies are ultraviolet <strong>and</strong> visible (UV-Vis),<br />

fluorescence <strong>and</strong> infrared (IR) spectroscopy. They differ only in the selection <strong>of</strong> the incident<br />

light wavelength. Both UV-Vis <strong>and</strong> fluorescence describe the phenomenon <strong>of</strong> electron<br />

72


excitation; namely, a valance electron <strong>of</strong> a molecule is excited upon absorbing energy from the<br />

electromagnetic radiation <strong>and</strong> is, thereby, transferred from one energy level to other more<br />

energetic level. The spectra are electronic. In contrast, IR spectra describe the vibration <strong>of</strong><br />

atoms (not electrons) around a chemical bond, as previously described in certain detail.<br />

An electron transition consists <strong>of</strong> the promotion <strong>of</strong> an electron from a molecular orbital in the<br />

ground state to an unoccupied orbital by absorption <strong>of</strong> a photon. The molecule is, then, said to<br />

be in an excited state. Let us recall first the various types <strong>of</strong> molecular orbitals. A orbital can<br />

be formed either from two s atomic orbitals, from one s <strong>and</strong> one p, or from two p atomic<br />

orbitals having a collinear symmetry axis. The bond formed in this way is called a bond. A<br />

orbital is formed from two p atomic orbitals overlapping laterally. The resulting bond is called a<br />

bond. For example, in ethylene (CH2=CH2) the two carbon atoms are linked by one <strong>and</strong> one<br />

bond. Absorption <strong>of</strong> appropriate energy can promote, for example, one <strong>of</strong> the electrons to<br />

an anti-bonding orbital denoted by . The transition is, then, called .<br />

A molecule may also possess non-bonding electrons located on heteroatoms such oxygen or<br />

nitrogen. The corresponding molecular orbitals are called orbitals. Promotion <strong>of</strong> a non-<br />

bonding electron to an anti-bonding orbital is possible, <strong>and</strong> the associated transition is<br />

denoted by . Hence, molecules containing a non-bonding electron, such as oxygen,<br />

nitrogen, sulphur, or halogens, <strong>of</strong>ten exhibit absorption in the UV region. To illustrate these<br />

energy levels, Figure 2.26 shows formaldehyde as an example with all their possible<br />

transitions. In particular, the transition deserves further attention: upon excitation, an<br />

electron is removed from the oxygen atom <strong>and</strong> goes into the orbital localized half on the<br />

carbon atom <strong>and</strong> half on the oxygen atom. The excited state, thus, has a charge<br />

transfer character, as shown by the increase observed in the dipole moment regarding the<br />

ground state dipole moment <strong>of</strong> C=O (40)(53).<br />

In summary, the electronic transitions observed in UV-Vis spectroscopy are ,<br />

, , , , <strong>and</strong> . The energy <strong>of</strong> these electronic transitions is,<br />

generally, in the following order: < < < < < . Of<br />

the six transitions outlined, only the two lowest energetic ones ( <strong>and</strong> ) are<br />

achieved by the energies available in the 200 to 800 nm. The last four types <strong>of</strong> electronic<br />

transitions required higher energy inputs, below 200 nm, corresponding to the far ultraviolet<br />

region <strong>of</strong> the electromagnetic spectrum. For instance, most transitions for individual<br />

bonds take place below 200 nm <strong>and</strong> a compound containing only bonds is transparent (near<br />

73


zero absorption) in the near-UV-Vis region. For example, methane (which has only C-H bonds)<br />

the transition takes place at 125 nm.<br />

In absorption <strong>and</strong> fluorescence spectroscopy, two important types <strong>of</strong> orbitals are considered:<br />

the Highest Occupied Molecular Orbitals (HOMO) <strong>and</strong> the Lowest Unoccupied Molecular<br />

Orbitals (LUMO). Both <strong>of</strong> these refer to the ground state <strong>of</strong> the molecule. For instance, in<br />

formaldehyde, the HOMO is the orbital <strong>and</strong> the LUMO is the orbital.<br />

Figure 2.26. Energy levels <strong>of</strong> molecular orbitals in formaldehyde <strong>and</strong> possible electronic<br />

transitions (39).<br />

When one <strong>of</strong> the two electrons <strong>of</strong> opposite spins (belonging to a molecular orbital <strong>of</strong> a<br />

molecule in the ground state) is promoted to a molecular orbital <strong>of</strong> higher energy, its spin is, in<br />

principle, unchanged so the total spin quantum number ( with<br />

zero. Because <strong>of</strong> the multiplicities <strong>of</strong> both the ground <strong>and</strong> excited states ( ) are<br />

equal to 1, both are called singlet states (usually denoted for the ground state) <strong>and</strong> , ,...<br />

for the excited states (see Figure 2.27). The corresponding transition is called a singlet-singlet<br />

transition. A molecule in a singlet excited state may undergo conversion into a state where the<br />

promoted electron has changed its spin; therefore, there are two electrons with parallel spins,<br />

<strong>and</strong> the total spin quantum number is 1 <strong>and</strong> the multiplicity is 3. Such state is called a triplet<br />

state because it corresponds to three states <strong>of</strong> equal energy. According to Hund’s rule, the<br />

triplet state has lower energy than the singlet state <strong>of</strong> the same configuration (39).<br />

or<br />

) is<br />

74


Figure 2.27. Distinction between singlet <strong>and</strong> triplet states.<br />

An instrument for measuring the absorption <strong>of</strong> UV <strong>and</strong> visible radiation is built <strong>of</strong> one or more<br />

light sources, a wavelength selector, a sample container, radiation transducers, a signal<br />

processors <strong>and</strong> readout devices. Figure 2.28 shows a scheme <strong>of</strong> a single-beam instrument for<br />

absorption measurements. Single beam instruments vary widely in their complexity <strong>and</strong><br />

performance characteristics. The simplest ones consist <strong>of</strong> a tungsten bulb as the source, a set<br />

glass filter for wavelength selection, a test tube for sample holders, a photovoltaic cell as the<br />

transducer, <strong>and</strong> an analogmeter as the readout device. Other sophisticated instruments have<br />

interchangeable tungsten <strong>and</strong> deuterium lamp sources, use a rectangular silica cell, <strong>and</strong> are<br />

equipped with a high-resolution grating monochromator with variable slit. Photomultiplier<br />

tubes are used as transducers, <strong>and</strong> the output is <strong>of</strong>ten digitalized, processed <strong>and</strong> stored in a<br />

computer so that it can be printer or plotted in several forms (54).<br />

Figure 2.28. General scheme <strong>of</strong> a single-beam instrument for UV-Vis absorption<br />

measurements.<br />

2.11.2.- Polarized light optical microscopy<br />

Polarized light optical microscopy (POM) provides all the benefits <strong>of</strong> bright field microscopy<br />

<strong>and</strong> yet <strong>of</strong>fers a wide range <strong>of</strong> information, which is simply not available with any other optical<br />

microscopy technique. POM can distinguish between isotropic <strong>and</strong> anisotropic materials. The<br />

technique exploits the optical properties <strong>of</strong> anisotropy to reveal detailed information about<br />

the structure <strong>and</strong> composition <strong>of</strong> materials.<br />

75


Isotropic materials demonstrate the same optical properties in all directions <strong>of</strong> space. They<br />

have only one refractive index <strong>and</strong> no restriction on the vibration direction <strong>of</strong> light passing<br />

through them. In contrast, anisotropic materials have optical properties that vary with the<br />

orientation <strong>of</strong> the incident light. They shown a range <strong>of</strong> refractive indexes depending both on<br />

the light propagation direction through the sample <strong>and</strong> on the vibrational plane coordinates<br />

(phenomenon known as birefringence). Birefringence is the term used to indicate the ability <strong>of</strong><br />

an anisotropic material to separate an incident ray into two rays, each polarized light ray<br />

travels at two different speeds. One <strong>of</strong> the resulting polarized rays travels through the sample<br />

with the same velocity in every direction <strong>and</strong> is termed the ordinary ray. The other ray travels<br />

with a velocity that is dependent upon the propagation direction within the anisotropic<br />

material. This light ray is termed the extraordinary ray. The two independent refractive indices<br />

<strong>of</strong> anisotropic sample are quantified in terms <strong>of</strong> their birefringence defined as: ,<br />

where <strong>and</strong> are the refractive indices <strong>of</strong> the extraordinary <strong>and</strong> ordinary rays, respectively.<br />

This expression is true for any part or fragment <strong>of</strong> anisotropic material except when light<br />

waves are propagated along the optical axis <strong>of</strong> the sample (55). The POM technique exploits<br />

the interference <strong>of</strong> the splitted light rays as they are re-united along the same optical path to<br />

extract information about these materials.<br />

Birefringent specimens exhibit characteristic patterns <strong>and</strong> orientations <strong>of</strong> light <strong>and</strong> dark<br />

contrast features that vary, depending on the shape <strong>and</strong> geometry <strong>of</strong> the object (linear or<br />

elongate vs. spherical) <strong>and</strong> the molecular orientation. In the absence <strong>of</strong> a compensator,<br />

spherical objects with radially symmetric molecular structure exhibit a dark upright<br />

polarization cross superimposed on a disk composed <strong>of</strong> four bright quadrants (55).<br />

POM is perhaps best known for its geological applications (primarily, for the study <strong>of</strong> minerals<br />

in rock-thin sections), but it can be used to study many other materials. These include<br />

minerals, composites, ceramics, fibers <strong>and</strong> polymers, <strong>and</strong> crystalline or highly ordered<br />

biological molecules such as DNA, protein, starch, wood <strong>and</strong> urea.<br />

As commented in detail previously, light can be represented as a transverse electromagnetic<br />

wave made up <strong>of</strong> mutually perpendicular, fluctuating electric <strong>and</strong> magnetic fields. It means<br />

that, at any point <strong>of</strong> a light beam, the magnetic field is always perpendicular to the electric<br />

field, <strong>and</strong> oscillates in a plane perpendicular to the propagating direction <strong>of</strong> the light beam. In<br />

plane polarized light, the electric field vector oscillates along a straight line <strong>of</strong> the propagation<br />

light beam. Figure 2.29 shows a polarized light optical microscope configuration. In this<br />

76


microscope, there are two polarized filters, the polarizer <strong>and</strong> the analyzer. The polarizer is<br />

situated below the specimen stage, <strong>and</strong> the analyzer is situated above the objectives <strong>and</strong> can<br />

be moved in <strong>and</strong> out <strong>of</strong> the light path as required. When both the analyzer <strong>and</strong> polarizer are in<br />

the optical path, in a crossed configuration, no light passes through the system <strong>and</strong> a dark field<br />

<strong>of</strong> view is present in the eyepieces.<br />

Figure 2.29. Polarized light optical microscope configuration.<br />

2.11.3.- Emission spectroscopy<br />

Inductively coupled plasma-atomic emission spectroscopy (ICP-AES) is a widespread technique<br />

for elemental analysis <strong>of</strong> a sample (56). In Figure 2.30, a scheme <strong>of</strong> an ICP-AES instrument is<br />

shown. A plasma source is used to make specific elements to emit light, after which a<br />

spectrometer separates this light in their characteristic wavelengths. This technique is<br />

especially suited for direct analysis <strong>of</strong> liquid samples. The sample solution is transformed into<br />

an aerosol by a nebuliser. Small droplets (1-10 μm) are transferred by an argon plasma. When<br />

the aerosol droplets enter the hot area <strong>of</strong> plasma, they are converted into salt particles. These<br />

salt particles are split into individual molecules that will subsequently fall apart to atoms <strong>and</strong><br />

ions. In the plasma, even more energy is transferred to the atoms <strong>and</strong> ions, promoting the<br />

excitation <strong>of</strong> their electrons. When these excited atoms <strong>and</strong> ions back to a lower excitation<br />

state or to their ground state they will emit electromagnetic radiation in the ultra-violet/visible<br />

range <strong>of</strong> the spectrum. The classical approach for ICP-AES measurement is to collect <strong>and</strong><br />

measure the emitted radiation radially, i.e. the optical axis is orthogonal to the central channel<br />

<strong>of</strong> the ICP. The limit <strong>of</strong> detection is in de order <strong>of</strong> ca. ng/ml.<br />

77


Figure 2.30. Scheme <strong>of</strong> an ICP-AES.<br />

2.11.4.-Superconducting quantum interference device (SQUID)<br />

SQUID is a sensitive device for measuring magnetic fields. SQUID magnetometers are used to<br />

characterize materials when the highest detection sensitivity over a broad temperatures range<br />

<strong>and</strong> application <strong>of</strong> magnetic fields up to several Tesla is needed. The main application <strong>of</strong> this<br />

instrument is the study <strong>of</strong> magnetic properties <strong>of</strong> materials by measuring the induced or<br />

remnant magnetic moment, usually as a function <strong>of</strong> applied magnetic field <strong>and</strong> temperature. A<br />

SQUID magnetometer combines several superconducting components, including the proper<br />

SQUID device, a superconducting magnet, detection coils, a flux transformer <strong>and</strong><br />

superconducting shields. To make a measurement, a sample is first attached to a sample rod.<br />

The sample is then scanned through the center <strong>of</strong> a first-order or second-order<br />

superconducting gradiometer. The gradiometers form a closed flux transformer that is coupled<br />

to a SQUID. The shape <strong>and</strong> magnitude <strong>of</strong> the response curve can be then analyzed using a<br />

computer to obtain a corresponding magnetic moment (57).<br />

Figure 2.31. Scheme <strong>of</strong> a SQUID magnetometer (57).<br />

78


In particular, we use SQUID to measure the magnetic properties <strong>of</strong> magnetic nanoparticles <strong>and</strong><br />

hybrids. In this regard, magnetic nanoparticles show a wide variety <strong>of</strong> unusual magnetic<br />

properties as compared to the respective bulk materials. In a paramagnetic material, the<br />

thermal energy overcomes the coupling forces between neighbouring atoms above the Curie<br />

temperature, causing r<strong>and</strong>om fluctuations in the magnetization directions that result in a null<br />

overall magnetic moment. However, in superparamagnetic materials the fluctuations affect<br />

the direction <strong>of</strong> individual entire crystallites. The magnetic moments <strong>of</strong> individual crystallites<br />

compensate each other <strong>and</strong> the overall magnetic moment becomes null. When an external<br />

magnetic field is applied, the behaviour is similar to paramagnetic materials except that,<br />

instead <strong>of</strong> each individual atom being independently influenced by an external magnetic field,<br />

the magnetic moment <strong>of</strong> entire crystallites aligns with the magnetic field (Figure 2.32) (58)(59).<br />

For magnetic nanoparticles, superparamagnetism occurs in magnetic materials composed <strong>of</strong><br />

very small crystallites (the threshold size depends on the nature <strong>of</strong> the material, i.e. Fe-based<br />

materials becomes superparamagnetic at sizes


increasing temperature after cooling the sample in the applied magnetic field (Figure 2.33). At<br />

temperatures well below the blocking temperature, the ZFC magnetization is small because<br />

the sample is not in thermal equilibrium, <strong>and</strong> the magnetization directions <strong>of</strong> the particles in a<br />

small applied field are mainly governed by the r<strong>and</strong>omly oriented directions <strong>of</strong> magnetization.<br />

With temperature increases, the smaller particles become superparamagnetic, <strong>and</strong> the<br />

probability <strong>of</strong> finding a particle with magnetization directions close to that <strong>of</strong> the applied field<br />

increases, resulting in an increase in the magnetization. With further temperature increases,<br />

more <strong>and</strong> more particles become superparamagnetic, resulting in an increasing magnetization<br />

until the net effect <strong>of</strong> the thermal energy becomes dominant <strong>and</strong> involves a decrease in the<br />

magnetization value. In the FC state, the magnetization in the blocked state is preferably<br />

frozen in a direction close to that <strong>of</strong> applied field. Therefore, the magnetization is considerably<br />

larger than that in the ZFC state. The ZFC <strong>and</strong> the FC curves coincide above the bifurcation<br />

temperature, which is the temperature above which all particles are superparamagnetic (59).<br />

Figure 2.33. ZFC <strong>and</strong> FC magnetization curves measured on a ferr<strong>of</strong>luid material (59).<br />

2.11.5.- Magnetic resonance imaging (MRI)<br />

One <strong>of</strong> the most valuable <strong>and</strong> widely used imaging methods in medical diagnostic is the<br />

Magnetic Resonance Image (MRI). Though the diagnostic potential <strong>of</strong> conventional MRI is<br />

immense <strong>and</strong> a variety <strong>of</strong> different contrast models are well established, further<br />

improvements <strong>of</strong> the method have been pursued in last years by application <strong>of</strong> magnetic<br />

nanoparticles for contrast enhancement based on functional site specificity.<br />

In MRI, the nuclear magnetic moment <strong>of</strong> protons is used as a sensitive probe <strong>of</strong> the chemical<br />

neighbourhood <strong>of</strong> protons in different tissues <strong>and</strong> organs <strong>of</strong> human body. Nuclear moments<br />

80


are aligned by means <strong>of</strong> an external magnetic bias field (commonly in the range <strong>of</strong> 0.2-3 T) <strong>and</strong><br />

the precession <strong>of</strong> the spins is excited by transverse RF pulses at a proton resonance frequency<br />

<strong>of</strong> about 42.58 MHz T -1 . After applying the pulse sequence, the induced magnetization decays<br />

<strong>and</strong> the longitudinal (T1) <strong>and</strong> transverse (T2) relaxation times <strong>of</strong> the precessing nuclear<br />

magnetic moments show tissue-specific differences that are used to generate the required<br />

image contrast. Imaging is performed by controlling external field gradients so that the<br />

resonance condition is fulfilled only in a restricted local region <strong>and</strong>, then, by scanning the<br />

resonant volume to be imaged. Magnetic response signals are detected by pick-up coils. In this<br />

way, the tissue-specific differences <strong>of</strong> the relaxation time T1 <strong>and</strong>/or T2 may be used for<br />

construction <strong>of</strong> the T1 <strong>and</strong> T2 weighted images showing optimal contrast <strong>of</strong> special tissue<br />

features. In practice, for optimization <strong>of</strong> tissue contrast a variety <strong>of</strong> different pulse sequences<br />

(e.g., the widely applied spin-echo methods) may be used (60)(61).<br />

Generally, a magnetic resonance scanner is essentially defined by three hardware groups <strong>and</strong><br />

their parameters: a) the main magnet with its homogeneity over imaging volume; b) the<br />

magnetic field gradient system with its linearity over the imaging volume; <strong>and</strong> c) the<br />

radi<strong>of</strong>requency (RF) system with its RF signal homogeneity <strong>and</strong> signal sensitivity over the<br />

imaging volume. Whole-body MRI imposes very special dem<strong>and</strong>s on these system components<br />

(62)(63):<br />

a) The main magnet. Magnetic resonance imaging requires a very strong magnetic field<br />

that has precisely the same magnitude <strong>and</strong> direction everywhere in the region we want to<br />

image. One <strong>of</strong> the key properties used to describe the quality <strong>of</strong> a MRI system is the<br />

uniformity, or homogeneity, <strong>of</strong> the applied magnetic field. For example, high-quality MRI<br />

systems made for clinical use in hospitals will have magnetic fields that vary less than 5 parts<br />

per million (ppm) over a 40 cm diameter spherical volume in the region desired for imaging.<br />

b) The magnetic field gradient. As mentioned earlier, a key property <strong>of</strong> the static<br />

magnetic field <strong>of</strong> a MRI system is its homogeneity, but anything we place inside the magnetic<br />

field tends to change the magnetic field slightly. To make the magnetic field as uniform as<br />

possible <strong>and</strong> to compensate for changes caused by different objects in the field, we shim the<br />

field. Shimming is typically h<strong>and</strong>led by placing small amounts <strong>of</strong> iron at specific locations within<br />

long trays that line the cylindrical magnetic field coil; or by several set <strong>of</strong> wire coils. Once the<br />

shimming is complete, the magnetic field is highly uniform over a central region where the<br />

imaging takes place.<br />

81


c) The radio frequency system. Radio frequency systems suitable for MRI are<br />

characterized by RF excitation with homogeneous signals over an examination volume as large<br />

as possible. Here, it is important that the excitation flip angle remains as constant as feasible,<br />

as the image contrast is fundamentally influenced by this parameter.<br />

Figure 2.34. A modern 9.4 Tesla MRI system (Bruker Biospin USR 94/20 instrument).<br />

2.12.- Laboratory equipment.<br />

Wilhelmy Plate method. Surface tensions () were determined by the Wilhelmy plate method<br />

using a Krüss K12 surface tension equipment, equipped with a processor to acquire the data<br />

automatically. The equipment was connected to a circulating water bath to keep the<br />

temperature constant to within 0.01 0 C.<br />

Drop tensiometer. Surface tension measurements were also carried out in a Track tensiometer<br />

equipment (I.T. Concept, France) adapted to determine surface tension values in real time with<br />

an accuaracy <strong>of</strong> 0.1 mN/m. Interfacial tension <strong>and</strong> interfacial rheology estimations are based<br />

82


on the digital pr<strong>of</strong>ile <strong>of</strong> a drop image <strong>and</strong> the resolution <strong>of</strong> the Gauss-Laplace equation. Win<br />

Drop s<strong>of</strong>tware (I.T. Concept, France) was used to obtain the surface tension values by means <strong>of</strong><br />

the axisymetric drop shape analysis. For air-water interface, an air-bubble upward was formed<br />

at the tip <strong>of</strong> a U-shaped stainless steel needle (0.5 mm inner diameter) immersed in water bulk<br />

phase. On the contrary, chlor<strong>of</strong>orm bubble downward was formed at the tip <strong>of</strong> a straight<br />

stainless steel needle (0.5 mm inner diameter).<br />

Langmuir-Blodgett trough. Surface-pressure isotherms were recorded in a Langmuir-Blodgett<br />

(LB) Teflon trough (model 611 from Nima Technologies, Ltc., Coventry, U.K.).<br />

Atomic force microscopy. The AFM measurements were performed in a JEOl instruments<br />

(model JSPM-4210) in non-contact mode using nitride cantilevers NSC 15 from Micro Masch,<br />

USA (typical working frequency <strong>and</strong> spring constant <strong>of</strong> 325 kHz <strong>and</strong> 40 N/m, respectively).<br />

83


Laser light scattering. Dynamic <strong>and</strong> static light scattering (DLS <strong>and</strong> (SLS) intensities were<br />

measured by means <strong>of</strong> an ALV-5000F (ALV-GmbH) instrument with vertically polarized incident<br />

light <strong>of</strong> wavelength λ= 488 nm supplied by a CW diode pumped Nd;YAG solid-state laser<br />

supplied by Coherent, Inc. <strong>and</strong> operated at 2 W.<br />

Polarized optical microscopy. POM experiments were made by using a Zeiss Axioplan optical<br />

microscope (Carl Zeiss Ltd., Welwyn Garden City, U.K.) A total magnification <strong>of</strong> either 50x or<br />

100 X was used. A polarizer <strong>and</strong> analyzer were put in fixed position, orthogonal to one another<br />

for the cross-polarizer imaging. Images were taken using a Kodak digital camera mounted on<br />

the top <strong>of</strong> the microscope.<br />

84


Laser confocal microscopy. Confocal microscopy was performed on a Leica TCS SP2 confocal<br />

system mounted on a Leica DM-IRE2 upright microscope, using a 40X objective. For<br />

fluorescence measurements, Thi<strong>of</strong>lavin T was added to spherulitic samples <strong>and</strong> excited using<br />

the 458 nm line <strong>of</strong> an argon ion laser. Simultaneously, transmission images were obtained with<br />

crossed polarizers in place using the 633 nm line <strong>of</strong> a He-Ne laser.<br />

Rheometry. The rheological properties <strong>of</strong> the samples were determined using a Bohlin CS10<br />

rheometer with water bath temperature control. Coutte geometry (bob, 24.5 mm in diameter,<br />

27 mm in height, cup, 26.5 in diameter, 29 mm in height) was used for fluid samples. Samples<br />

<strong>of</strong> very high modulus were investigated using cone <strong>and</strong> plate geometry (diameter 40 mm,<br />

angle 4 0 ). Frequency scans <strong>of</strong> storage (G’) <strong>and</strong> lost (G’’) moduli were recorded for selected<br />

copolymer concentrations <strong>and</strong> temperatures with the instrument in oscillatory shear mode<br />

<strong>and</strong> with the strain amplitude (A) maintained at a low value (A < 0.5 %) by means <strong>of</strong> auto-<br />

stress facility <strong>of</strong> the Bohlin s<strong>of</strong>tware. This ensured that measurements <strong>of</strong> G’ <strong>and</strong> G’’ were in the<br />

linear viscoelastic region.<br />

85


Transmission electron microscopy. TEM micrographs were acquired by using a Phillips CM-12<br />

electron microscope operating at an accelerating voltage <strong>of</strong> 120 kV.<br />

High resolution transmission electron microscopy <strong>and</strong> selected area electron diffraction. HR-<br />

TEM images <strong>and</strong> SAED patterns were obtained with a transmission electron microscope (Carl-<br />

Zeiss Libra 200 FE-EFTEM) operating at 200 kV.<br />

Scanning electron microscopy. SEM micrographs were acquired by using a LEO-435 VP<br />

scanning electron microscope (Leica Microsystems Gmb H, Wetlar, Germany) operating at an<br />

accelerating voltage <strong>of</strong> 30 kV. Microanalysis <strong>of</strong> the scanning electron microscopy samples was<br />

performed.<br />

86


Fluorescence spectroscopy. Fluorescence was measured in a Cary Eclipse fluorescence<br />

spectrophotometer equipped with a temperature control device <strong>and</strong> a multicell sample holder<br />

(Varian instrument Inc.).<br />

Ultraviolet-visible spectroscopy. UV-vis data were acquired by using a spectrophotometer DU<br />

series 640, Beckman Coulter, (Fullerton, CA) operating at 190-110 nm.<br />

Far <strong>and</strong> near-ultraviolet circular dichroism spectroscopy. CD spectra were obtained using a<br />

JASCO-715 automatic recording spectropolarimeter (Jasco, Tokyo, Japan) with a JASCO PTC-<br />

343 Peltier-type thermostated cell holder. Quartz cuvettes with 0.2cm pathlength were used.<br />

CD spectra were recorded between 195 <strong>and</strong> 300 nm at 25 0 C.<br />

87


Fourier transform infrared spectroscopy. An FTIR spectrometer (model IFS-66V; Bruker)<br />

equipped with a horizontal ZnS ATR accessory. The spectra were obtained at a resolution <strong>of</strong> 2<br />

cm -1 <strong>and</strong> 200 scans were accumulated to obtain a reasonable signal/noise ratio.<br />

X-ray diffraction. XRD experiments were carried out using a Siemens D5005 rotating anode X-<br />

ray generator. Twin Göbel mirror were used to produce a well-collimated beam <strong>of</strong> Cu Kα<br />

radiation (λ = 1.5418 Å). XRD patterns were recorded with an imaging plate detector AXS F.Nr.<br />

J2-394.<br />

88


Superconducting quantum interference device. Magnetic susceptibility measurements were<br />

carried out with a SQUID magnetometer (Quantum Design MPMS5, San Diego, CA).<br />

Magnetic resonance measurements. Transverse <strong>and</strong> longitudinal relaxation times were<br />

measured at 9.4 T (400 MHz) with a Brucker Biospin USR 94/20 instrument (Ettlingen,<br />

Germany).<br />

89


Contents<br />

3.1<br />

3.2<br />

3.3<br />

3.4<br />

3.5<br />

3.1.- Introduction<br />

Introduction<br />

Behaviour <strong>of</strong> block copolymers in solution<br />

CHAPTER 3<br />

Block copolymers<br />

Mechanisms <strong>of</strong> morphologic control <strong>and</strong> thermodynamic<br />

stability<br />

Block copolymers <strong>of</strong> polyoxyalkylenes in aqueous solution<br />

E12S10, E10S10E10 <strong>and</strong> E137S18E137 block copolymers<br />

Block copolymers are a particular class <strong>of</strong> polymers that are included in a wide family <strong>of</strong><br />

nanomaterials, known as s<strong>of</strong>t materials. Independently <strong>of</strong> the synthetic procedure, block<br />

copolymers can simply be considered as being formed by two or more chemically<br />

homogeneous polymer fragments joined together by covalent bonds, which can be found in<br />

different positions along the polymeric chain (Figure 3.01) (64)(65). According to the number<br />

<strong>of</strong> the building units, these polymers can be denominated as diblock, triblock <strong>and</strong> multiblock<br />

copolymers. The molecular architecture <strong>of</strong> block copolymers can be as simple as a linear chain<br />

or more complex like the molecular architecture <strong>of</strong> star or graft copolymers, which can be<br />

controlled by means <strong>of</strong> the chemical synthetic procedures (9)(66)-(68).<br />

Methodologies used in the synthesis <strong>of</strong> block copolymers allow designing any polymeric chain<br />

with specific properties <strong>and</strong> architectures. Generally, the structure <strong>of</strong> a block copolymer is<br />

composed <strong>of</strong> incompatible joint homopolymers. When a copolymer is molten, <strong>and</strong><br />

subsequently left cool down slowly, the incompatibilities between constituent building blocks<br />

give rise to microphase separation, leading to the formation <strong>of</strong> ordered nanodomains. The<br />

covalent linkage between the different polymeric chains prevents the separation in<br />

macrodomains. The generation <strong>of</strong> these structural nanodomains is managed by the self-<br />

assembling process <strong>of</strong> homologous segments that compose the block copolymer. Also, such<br />

dimensional nanostructures depend on the volume fraction <strong>of</strong> blocks copolymer. Two<br />

competing effects govern the thermodynamics <strong>of</strong> block copolymer melts. At high<br />

91<br />

93<br />

95<br />

96<br />

97<br />

91


temperatures, the chains are mixed homogeneously, as in any polymer melt. As the<br />

temperature is reduced the tendency for the blocks to segregate is enhanced, i.e. the enthalpic<br />

process <strong>of</strong> demixing is favoured. However, this is necessary accompanied by a reduction in the<br />

entropy as the chain configuration becomes more constrained. The extent <strong>of</strong> segregation <strong>of</strong><br />

the copolymer may be expressed using the reduced parameter N. Here, is the Flory-Huggins<br />

interaction parameter, which contains a significant enthalpic contribution, <strong>and</strong> is governed by<br />

the incompatibility <strong>of</strong> the monomers; <strong>and</strong> N is the copolymer degree <strong>of</strong> polymerization,<br />

reflecting the N-dependent translational <strong>and</strong> configurational entropy. The transition from<br />

homogeneous melt <strong>of</strong> chains to heterogeneous melt <strong>of</strong> ordered microphase-separated<br />

domains occurs at a critical value <strong>of</strong> N, depending on the composition <strong>of</strong> the copolymer (65).<br />

In bulk, the minority block is segregated from the majority block forming regularly shaped <strong>and</strong><br />

uniformly-spaced nanodomains. The shape <strong>of</strong> the segregated domains in a diblock is governed<br />

by the volume fraction <strong>of</strong> the minority block, f, <strong>and</strong> block incompability. Figure 3.1 shows the<br />

equilibrium morphologies documented for diblock copolymers as an example. At a volume<br />

fraction <strong>of</strong> ≈ 20 %, the minority block forms a body-centred cubic spherical phase in the matrix<br />

<strong>of</strong> the majority block. It changes to hexagonally packed cylinders at a volume fraction ≈ 30 %.<br />

Alternating lamellae are formed at approximately equal volume fractions for the two blocks. At<br />

a volume fraction <strong>of</strong> ≈ 38 %, the minority block forms bicontinuous structures at moderate <strong>and</strong><br />

high incompatibility, respectively (64)(65)(69).<br />

A Block B Block<br />

Sphere Cylindric Bicontinuous Lamellar<br />

Figure 3.1. Regular structures in block copolymer leaded by theirs self-assembling.<br />

The physical properties <strong>of</strong> copolymeric materials have been applied in a wide range <strong>of</strong> fields.<br />

For example, by decades copolymeric materials in solid state have been used as thermoplastic<br />

<strong>and</strong> elastomeric materials, foams, <strong>and</strong> glues owing to their high compatibility, sensitivity, <strong>and</strong><br />

92


esistance to external physical factors (65). On other h<strong>and</strong>, the auto-association properties <strong>of</strong><br />

block copolymers in selective media allow the formation <strong>of</strong> structure-ordered aggregates, <strong>and</strong>,<br />

for that reason, this type <strong>of</strong> polymeric materials has been received special attention in last<br />

years for their used in high-nanotechnology applications, as shown below (8)(9).<br />

3.2.- Behaviour <strong>of</strong> block copolymers in solution.<br />

The hydrophilic <strong>and</strong> hydrophobic nature <strong>of</strong> amphiphilic block copolymer molecules lead to<br />

their self-assembly into a variety <strong>of</strong> structures in solution. There are two basic processes that<br />

characterize the phase behaviour <strong>of</strong> block copolymers in solution: micellization <strong>and</strong> gelation.<br />

Micellization occurs in dilute solution <strong>of</strong> block copolymers in a selective solvent at a fixed<br />

temperature above the cmc, as commented in more detail below. At higher copolymer<br />

concentrations, the micelles can order onto a lattice above a critical gel concentration (cgc).<br />

These regimes are illustrated in Figure 3.2.<br />

Figure 3.2 Illustration <strong>of</strong> cmc <strong>and</strong> cgc in ablock copolymer solution.<br />

When a block copolymer is dissolved in a selective solvent, that is, a solvent which is good for<br />

one <strong>of</strong> the blocks <strong>and</strong> bad for the other one, the lyophobic part <strong>of</strong> the macromolecule tends to<br />

segregate <strong>and</strong> auto-associate with their neighbouring molecules, leading to the formation <strong>of</strong><br />

stable aggregates with specific structures, in order to avoid the direct contact between solvent<br />

molecules with the insoluble copolymer block (70). The spontaneous aggregation <strong>of</strong> block<br />

copolymers is carried out either when the concentration <strong>of</strong> the macromolecule is near to their<br />

critical micellar concentration, or on the other h<strong>and</strong>, by changes in the solution temperature.<br />

In this case, the critical micellar temperature is reached. Both parameters are fundamental to<br />

characterize the association properties <strong>of</strong> a copolymer-solvent system (71). Despite the cmc<br />

does not correspond to a thermodynamic property <strong>of</strong> the system, it can simply be defined<br />

phenomenollogically as the concentration at which a sufficient number <strong>of</strong> micelles is formed to<br />

93


e detected by a given method. Thermodynamically, for copolymers with a narrow block-<br />

length, the association process corresponds to equilibrium between molecules (unimers), A,<br />

<strong>and</strong> micelles, AN, containing N molecules. This can be written as:<br />

If the association number is large <strong>and</strong> independent <strong>of</strong> the temperature then <strong>and</strong><br />

the st<strong>and</strong>art Gibbs energy <strong>of</strong> association is:<br />

Under this condition, for molecules <strong>and</strong> micelles in equilibrium just above the cmc<br />

Figure 3.3. Schematic representation <strong>of</strong> block copolymer micelles in aqueous solution, dense<br />

hydrophobic core is constituted by the hydrophobic block, while the micellar shell is formed by<br />

the hydrophilic block.<br />

The micellization process <strong>of</strong> amphiphilic block copolymers has been studied in a wide range <strong>of</strong><br />

selective solvents: from water to organic polar solvents <strong>and</strong> non-polar solvents, as well as<br />

ecological solvents like supercritic fluids <strong>and</strong> ionic liquids (72). The simplest structural<br />

aggregates formed by block copolymer are called micelles, by the analogy that exists with the<br />

structure <strong>of</strong> common micelle formed by the convectional surfactants in aqueous media (Figure<br />

3.3) (73).<br />

3.1<br />

3.2<br />

94


In aqueous media, the micelle structure consists <strong>of</strong> a dense hydrophobic core surrounded by a<br />

hydrophilic shell which is exposed toward the bulk solvent phase in a spherical configuration;<br />

the micelles can change their size from 5 to some hundreds nanometers <strong>and</strong> can possess<br />

different geometries (spheres, worn-like, <strong>and</strong> so on) (74)-(78).<br />

The various reported micellar-like morphologies are primarily a result <strong>of</strong> the inherent<br />

molecular curvature <strong>and</strong> how this influences the packing <strong>of</strong> the copolymer chains: specific self-<br />

assembled nanostructures can be targeted according to a dimensionless packing parameter, ,<br />

which is defined as<br />

, where is the volume <strong>of</strong> the hydrophobic chains, is the<br />

optimal area <strong>of</strong> the head group, <strong>and</strong>, is the legth <strong>of</strong> the hydrophobic tail, as occurred for<br />

classical surfactants. Therefore, the packing parameter <strong>of</strong> a given molecule usually dictates its<br />

most likely self-assembled morphology. As a general rule, spherical polymeric micelles are<br />

favored when<br />

, cylindrical micelles when<br />

structures (vesicles, also known as polymersomes) when<br />

, <strong>and</strong> enclosed membrane<br />

Despite the structured aggregates from block copolymers formed in aqueous solution are<br />

similar to those produced by the self-assembly <strong>of</strong> small surfactant molecules such as lipids <strong>and</strong><br />

detergents, the polymeric nature <strong>of</strong> the amphiphilic macromolecules provides new special<br />

properties to these structural complexes, as stability <strong>and</strong> functionality. In this way, their<br />

potential applications in different technological areas are largely increasing, including<br />

materials science, bioengineering, medicine, pharmaceutics industry, <strong>and</strong> ink <strong>and</strong> paint<br />

industry (70)(79).<br />

3.3.- Mechanism <strong>of</strong> morphologic control <strong>and</strong> thermodynamic stability.<br />

The conformation <strong>of</strong> the block copolymer chains depends on the interaction between chain<br />

segments <strong>and</strong> solvent molecules. In aqueous solution, the structure <strong>of</strong> the aggregates is<br />

governed by the balance <strong>of</strong> the interaction forces between the block copolymer molecules<br />

with water molecules, <strong>and</strong> by the intrachain interactions between the blocks which constitute<br />

the copolymer molecule (where the hydrophilic chain exp<strong>and</strong>s, avoiding unfavourable<br />

segment-segment interactions <strong>and</strong> the hydrophobic part contract to minimize the contact with<br />

solvent molecules).<br />

It is possible to achieve a precise control <strong>of</strong> the size <strong>and</strong> shape <strong>of</strong> polymeric aggregates, <strong>and</strong><br />

.<br />

95


subsequently, <strong>of</strong> their thermodynamic stability by modification <strong>of</strong> some factors such as block<br />

copolymer molecular weight, the relatively length <strong>of</strong> both hydrophobic <strong>and</strong> hydrophilic blocks,<br />

the chemical nature <strong>of</strong> the different monomer units <strong>and</strong> the architecture <strong>of</strong> the polymeric<br />

backbone. Firstly, this can be achieved by means <strong>of</strong> the block copolymer synthetic processes.<br />

However, the synthesis <strong>of</strong> block copolymers is time-consuming <strong>and</strong> expensive, disadvantages<br />

that have to be taken into account. A simpler alternative is the possibility <strong>of</strong> manipulating the<br />

block copolymer solution properties through the change <strong>of</strong> external parameters such as<br />

temperature, salinity, pH, addition <strong>of</strong> cosolvents… Under this context, it is possible to obtain<br />

different aggregate sizes or morphologies using the same block copolymer or the same family<br />

<strong>of</strong> copolymers. The morphology <strong>of</strong> the structured aggregates can be determined by a balance<br />

<strong>of</strong> different factors such as the solvent properties, the conformation <strong>of</strong> the hydrophobic chains<br />

into the micelle core, the repulsion between hydrophilic blocks into the micelle shell, <strong>and</strong> the<br />

surface tension between the micelle cores <strong>and</strong> the bulk solution (77)(78).<br />

3.4.- Block copolymers <strong>of</strong> polyoxyalkylenes in aqueous solution.<br />

The block copolymers <strong>of</strong> polyoxyalkylenes are composed <strong>of</strong> a hydrophilic block constituted by<br />

poly(ethylene oxide) (OCH2CH2 ethylene oxide unit, E) <strong>and</strong> by a second hydrophobic block that<br />

can be made <strong>of</strong>, for example, poly(propylene oxide) (P), poly(styrene oxide) (S), or<br />

poly(butylene oxide) (B), amongst others. In aqueous medium, polyoxyalkylene amphiphilic<br />

block copolymer contains both the hydrophilic <strong>and</strong> hydrophobic blocks that can associate to<br />

usually form spherical micelles in dilute aqueous solution. Increasing the concentration leads<br />

to the formation <strong>of</strong> lyotropic crystal mesophases (gels) where the structural objects are<br />

micelles, usually efficiently packed in a body-centered or face-centered cubic structure as a<br />

result <strong>of</strong> their interaction as hard spheres (80)(81).<br />

For poly(oxyalkylenes) block copolymers, values <strong>of</strong> the cmc are affected by the variation <strong>of</strong><br />

hydrophilic polyoxyethylene <strong>and</strong> hydrophobic block lengths. It is known that the dependence<br />

<strong>of</strong> the cmc on the number <strong>of</strong> E units in a copolymer is small unless the hydrophobic-block<br />

length is short. Indeed, when comparing results for two copolymers <strong>of</strong> nominally the same<br />

hydrophobic-block length, the effect <strong>of</strong> E-block length may well be less than the variation<br />

caused by the uncertainty in hydrophobic-block length. However, within a series <strong>of</strong> copolymers<br />

it is desirable to account for variation in E-block length. Taking as an example PE-PP-PE<br />

Pluronic block copolymers, micelle formation <strong>of</strong> this class <strong>of</strong> polymers in water is believed to<br />

96


esult from dehydratation <strong>of</strong> the hydrophobic block with increasing temperature, water being<br />

a selective solvent for PE. This explains why the solution properties <strong>of</strong> Pluronic polymers are<br />

strongly temperature-dependent. For most PE-PP-PE copolymers, the cmt increases with<br />

decreasing in copolymer concentration. The number <strong>of</strong> oxypropylene units has a strong effect<br />

on the micellization, copolymers with large hydrophobic PP block length forming micelles at<br />

much lower concentrations. Indeed, the cmc is found to decrease exponentially with the PP<br />

block length. On the other h<strong>and</strong>, the PE block has a smaller influence on the micellization. An<br />

increase in the number <strong>of</strong> PE units leads to small increases in the cmc <strong>and</strong> cmt values. The cmt<br />

<strong>and</strong> cmc decrease with increasing the total copolymer molecular weight, if comparisons are<br />

made for a constant PP/PE ratio. On the other h<strong>and</strong>, the large positive enthalpy <strong>of</strong><br />

micellization, <strong>and</strong> the st<strong>and</strong>ard free energy <strong>of</strong> micellization observed on the self-assembly <strong>of</strong><br />

PE-PP-PE demonstrate that the association process is entropy driven. Chain architecture has<br />

also a great influence on micelle formation. For example, the tendency for micellization <strong>of</strong> PP-<br />

PE-PP copolymer is reduced compared to a PE-PP-PE copolymer <strong>of</strong> the same composition.<br />

3.5.- E12S10, E10S10E10 <strong>and</strong> E137S18E137 block copolymers<br />

Our research group has been collaborating during last years with research groups <strong>of</strong> the School<br />

<strong>of</strong> Pharmacy <strong>and</strong> Pharmaceutical Sciences <strong>and</strong> the School <strong>of</strong> Chemistry <strong>of</strong> the University <strong>of</strong><br />

Manchester in synthesizing block copolymers formed by ethylene oxide <strong>and</strong> styrene oxide<br />

blocks (81)-(83). In this regard, it has been previously demonstrated that the micellization<br />

process <strong>of</strong> diblock copolymers <strong>of</strong> this type has not cmt values. Moreover, they usually have<br />

small cmc values <strong>and</strong> micellization enthalpies (micH) as a consequence <strong>of</strong> the high<br />

hydrophobicity <strong>of</strong> the styrene oxide block. Triblock copolymers show cmc values higher than<br />

diblock ones for the same hydrophobic block length. On the other h<strong>and</strong>, the cmc values <strong>of</strong> the<br />

triblocks weakly depend <strong>of</strong> the hydrophilic block length.<br />

Block copolymers studied in this work were: E12S10, E10S10E10 <strong>and</strong> E137S18E137, (where E<br />

represents the oxyethylene unit, S represents the oxystyrene unit, <strong>and</strong> the subscripts<br />

correspond to block lengths). These copolymers were synthesised by sequential anionic<br />

polymerisation <strong>of</strong> the two monomers, oxyethylene <strong>and</strong> oxystyrene. Oxyrane compounds are<br />

heterocyclic compounds constructed by two carbon atoms <strong>and</strong> one oxygen atom into the<br />

cyclic structure (Figure 3.4) (84).<br />

97


a) O<br />

b)<br />

H 2C CH 2<br />

O<br />

H 2C CH<br />

Figure 3.4. Oxyranes used in the synthesis <strong>of</strong> the block copolymers a) ethylene oxide <strong>and</strong> b)<br />

styrene oxide.<br />

Anionic polymerization proceeds via organometallic sites, carbanions (or oxanions) with<br />

metallic counterions. Carbanions are nucleophiles; consequently, the monomers that can be<br />

polymerized by anionic polymerization are those bearing an electro-attractive substituent on<br />

the polymerizable double bound. Initiation <strong>of</strong> polymerization is accomplished by analogous<br />

low molecular weight organometallic compounds (initiators). A wide variety <strong>of</strong> initiators has<br />

been used so far in order to produce living polymers. Among them, the most widely used are<br />

organolithiums. The main requirement for employment <strong>of</strong> an organometallic compound as an<br />

anionic initiator is its rapid reaction with the monomer at the initiation step <strong>of</strong> the<br />

polymerization reaction <strong>and</strong>, specifically, with a reaction rate larger than that <strong>of</strong> the<br />

propagation step. This leads to the formation <strong>of</strong> polymers with narrow molecular weight<br />

distributions because all active sites start polymerizing the monomer almost at the same time.<br />

Propagation proceeds through nucleophilic attack <strong>of</strong> the carbanionic site onto a monomer<br />

molecule with reformation <strong>of</strong> the first anionic active center. The situation is similar in the case<br />

<strong>of</strong> the ring opening polymerization <strong>of</strong> cyclic monomers containing heteroatoms (oxyranes,<br />

lactones, siloxanes, etc.). Under appropriate experimental conditions, due to the absence <strong>of</strong><br />

termination <strong>and</strong> chain transfer reactions, carbanions (or, in general, anionic sites) remain<br />

active after complete consumption <strong>of</strong> monomers, giving the possibility <strong>of</strong> block copolymer<br />

formation, in the simplest case, by introduction <strong>of</strong> a second monomer into the polymerization<br />

mixture. However, a variety <strong>of</strong> different synthetic strategies have been reported for the<br />

preparation <strong>of</strong> linear block copolymers by anionic polymerization.<br />

As an example, in the polymerisation reaction <strong>of</strong> a mon<strong>of</strong>unctional poly(oxyethylene) the<br />

addition <strong>of</strong> ethylene oxide (E) is initiated by a mon<strong>of</strong>unctional alcohol (ROH), part <strong>of</strong> which is in<br />

the form <strong>of</strong> its alkali-metal salt (e.g., RO-K + .) Initiation is instantaneous, <strong>and</strong> the subsequent<br />

reaction scheme is written:<br />

98


where Em represents a growing chain, the active centre is an ion pair, <strong>and</strong> the rapid<br />

equilibration ensures that all chains are equally likely to grow. This ideal scheme would<br />

produce polymers with Mn= (mass <strong>of</strong> E polymerised)/(moles <strong>of</strong> initiator used) <strong>and</strong> a Poisson<br />

distribution <strong>of</strong> chain lengths. In practice, the chain length distribution can be widened by a<br />

slow initiation. In the absence <strong>of</strong> termination, the system is living <strong>and</strong> is ideal for preparation<br />

<strong>of</strong> block copolymers. The same scheme applies to styrene oxide.<br />

It is known that block poly(oxyalkylene) copolymers with hydrophobic blocks formed by<br />

propylene oxide or butylene oxide (either diblock or triblock architectures) have a<br />

spontaneous formation <strong>of</strong> spherical <strong>and</strong> cylindrical micelles in dilute aqueous solution. On the<br />

other h<strong>and</strong>, formation <strong>of</strong> vesicular structures for poly(oxyalkylene) block copolymers needs<br />

<strong>of</strong>tenly an energy supply via mechanical, sonochemical or electrical methods (85)(86).<br />

Nevertheless, it has been previously reported the spontaneous formation <strong>of</strong> vesicles (87)(88),<br />

where these types <strong>of</strong> structures are likely to form because <strong>of</strong> relatively large volume to length<br />

ratios <strong>of</strong> the hydrophobic block <strong>and</strong> the relative lengths <strong>of</strong> the poly(oxyethylene) blocks. In the<br />

case <strong>of</strong> block copolymers containing poly(oxystyrene) blocks as the hydrophobic block, the<br />

formation <strong>of</strong> elongated micelles has been unusual, whereas vesicle formation has not been yet<br />

reported.<br />

99


Contents<br />

4.1<br />

4.2<br />

4.3<br />

4.3.1<br />

CHAPTER 4<br />

Papers on block copolymers<br />

<strong>Self</strong>-<strong>Assembly</strong> process <strong>of</strong> different poly(oxystyrene)-<br />

poly(oxyethylene) block copolymers: spontaneous<br />

formation <strong>of</strong> vesicular structures <strong>and</strong> elongated micelles.<br />

Surface properties <strong>of</strong> monolayers <strong>of</strong> amphiphilic<br />

poly(ethylene oxide)-poly(styrene oxide) block copolymers.<br />

Appended papers<br />

Relevant aspects<br />

97<br />

107<br />

117<br />

118


<strong>Self</strong>-<strong>Assembly</strong> Process <strong>of</strong> Different Poly(oxystyrene)-Poly(oxyethylene)<br />

Block Copolymers: Spontaneous Formation <strong>of</strong> Vesicular Structures<br />

<strong>and</strong> Elongated Micelles<br />

Josué Juárez, † Pablo Taboada,* ,† Miguel A. Valdez, ‡ <strong>and</strong> Víctor Mosquera †<br />

Departamento de InVestigación en Polímeros y Materiales y Departamento de Física, UniVersidad de<br />

Sonora, Sorales Resales y TransVersal, 83000 Hermosillo Sonora, Mexico, <strong>and</strong> Laboratorio de Física de<br />

Coloides y Polímeros, Grupo de Sistemas Complejos, Departamento de Física de la Materia Condensada,<br />

Facultad de Física, UniVersidad de Santiago de Compostela, Spain<br />

ReceiVed February 12, 2008. ReVised Manuscript ReceiVed March 26, 2008<br />

In the present work, we investigated the micellization, gelation, <strong>and</strong> structure <strong>of</strong> the aggregates <strong>of</strong> three poly(ethylene<br />

oxide)-polystyrene oxide block copolymers (E12S10, E10S10E10, <strong>and</strong> E137S18E137, where E denotes ethylene oxide <strong>and</strong><br />

S styrene oxide <strong>and</strong> the subscripts the block length) in solution. Two <strong>of</strong> them have similar block lengths but different<br />

structures (E12S10 <strong>and</strong> E10S10E10) <strong>and</strong> the other has longer blocks (E137S18E137). For the first time, the spontaneous<br />

formation <strong>of</strong> vesicles by a poly(oxystyrene)-poly(oxyethylene) block copolymer is reported. These vesicular structures<br />

are present when copolymer E12S10 self-assembles in aqueous solution in coexistence with spherical micelles, as<br />

confirmed by the size distribution obtained by dynamic light scattering <strong>and</strong> pictures obtained by polarized optical<br />

microscopy, <strong>and</strong> transmission <strong>and</strong> cryo-scanning electron microscopies. Vesicle sizes vary between 60 <strong>and</strong> 500 nm.<br />

On the other h<strong>and</strong>, for copolymers E10S10E10 <strong>and</strong> E137S18E137, only one species is found in solution, which is assigned<br />

to elongated <strong>and</strong> spherical micelles, respectively. If we compare the high aggregation number derived by static light<br />

scattering for the triblock block copolymer micelles, with the maximum theoretical micellar dimensions compatible<br />

with a spherical geometry, we can see that the micellar geometry cannot be spherical but must be elongated. This<br />

is corroborated by transmission electron microscopy images. On the other h<strong>and</strong>, tube inversion was used to define<br />

the mobile-immobile (s<strong>of</strong>t-hard gel) phase boundaries. To refine the phase diagram <strong>and</strong> observe the existence <strong>of</strong><br />

additional phases, rheological measurements <strong>of</strong> copolymer E137S18E137 were done. The results are in good agreement<br />

with previous values published for other polystyrene oxide-poly(ethylene oxide) block copolymers. In contrast, copolymers<br />

E12S10 <strong>and</strong> E10S10E10 did not gel in the concentration range analyzed. Thus, only certain concentrations <strong>of</strong> copolymer<br />

E10S10E10 were analyzed by rheometry, for which an upturn in the low-frequency range <strong>of</strong> the stress moduli was<br />

observed, denoting an evidence <strong>of</strong> an emerging slow process, which we assign to the first stages <strong>of</strong> formation <strong>of</strong> an<br />

elastic network.<br />

Introduction<br />

Block copolymers <strong>of</strong> hydrophilic poly(oxyethylene)<br />

(OCH2CH2, E) oxyethylene unit) <strong>and</strong> hydrophobic poly(oxystyrene)<br />

(OCH2CH(C6H5), S ) oxyphenylethylene unit) are<br />

denoted as either EmSn,SnEmor EmSnEm,SnEmSn, depending on<br />

the sequence <strong>of</strong> polymerization or architecture, respectively. They<br />

associate to usually form spherical micelles in dilute aqueous<br />

solution. 1,2 Increases in concentration lead to the formation <strong>of</strong><br />

lyotropic crystal mesophases (gels) where the structural objects<br />

are the micelles usually efficiently packed in a body-centered or<br />

face-centered cubic structure as a result <strong>of</strong> being effectively<br />

interacting as hard spheres. 3<br />

Aggregate morphology is determined primarily by a force<br />

balance among three contributions: the core-chain stretching,<br />

corona-chain repulsion, <strong>and</strong> interfacial tension between the core<br />

<strong>and</strong> the outside solution. Consequently, factors that influence the<br />

above balance can be used to control the aggregate architecture:<br />

* Author to whom correspondence should be addressed. Telephone:<br />

0034981563100, ext.: 14042. Fax: 0034981520676. E-mail: fmpablo@usc.es.<br />

† Universidad de Santiago de Compostela.<br />

‡ Universidad de Sonora.<br />

(1) Crothers, M.; Attwood, D.; Collett, J. H.; Yang, Z.; Booth, C.; Taboada,<br />

P.; Mosquera, V.; Ricardo, N. P. S.; Martini, L. G. A. Langmuir 2002, 18, 8685–<br />

8691.<br />

(2) Yang, Z.; Crothers, M.; Ricardo, N. M. P. S.; Chaibundit, C.; Taboada,<br />

P.; Mosquera, V.; Kelarakis, A.; Havredaki, V.; Martini, L.; Valder, C.; Collett,<br />

J. H.; Attwood, D.; Heatley, F.; Booth, C. Langmuir 2003, 19, 943–950.<br />

(3) Castelletto, V.; Hamley, I. W.; Crothers, M.; Attwood, D.; Yang, Z.; Booth,<br />

C. J. Macromol. Sci., Phys. 2004, 43, 13–27.<br />

Langmuir 2008, 24, 7107-7116<br />

10.1021/la8004568 CCC: $40.75 © 2008 American Chemical Society<br />

Published on Web 06/12/2008<br />

7107<br />

polymer molecular weight, relative block lengths, hydrophobic<br />

block structure, solution conditions, <strong>and</strong> so forth. For a given<br />

value <strong>of</strong> the association number (Nw) for a poly(oxyalkylene)<br />

block copolymer, the average volume <strong>of</strong> a micelle core is readily<br />

estimated, <strong>and</strong> the corresponding spherical radius (rc) can be<br />

compared with the length l (or half-length, l/2, in the case <strong>of</strong><br />

triblock copolymers) <strong>of</strong> the extended hydrophobic blocks. With<br />

allowance made for a Poisson distribution <strong>of</strong> hydrophobic blocks,<br />

experience shows that spherical (or near-spherical) micelles will<br />

be formed if l (l/2 for triblock architecture) is approximately<br />

equal or greater than the value <strong>of</strong> rc. 4,5 If this does not occur,<br />

other types <strong>of</strong> assemblies are observed under certain solution<br />

conditions. For example, a cylindrical geometry is expected as<br />

the solution temperature is increased because <strong>of</strong> an increase in<br />

the aggregation number derived from block copolymer dehydration:<br />

6 the core blocks are stretched as the radius <strong>of</strong> a spherical<br />

micelle core is increased eventually to reach a point at which the<br />

enthalpy penalty forces the transition to a cylindrical geometry.<br />

Thus, formation <strong>of</strong> cylindrical micelles has been widely studied<br />

in block poly(oxyalkylene) copolymers with hydrophobic blocks<br />

(4) Taboada, P.; Velazquez, G.; Barbosa, S.; Castelletto, V.; Nixon, S. K.;<br />

Yang, Z.; Heatley, F.; Hamley, I. W.; Ashford, M.; Mosquera, V.; Attwood, D.;<br />

Booth, C. Langmuir 2005, 21, 5263–5271.<br />

(5) Ricardo, N. M. P. S.; Pinho, M. E. N.; Yang, Z.; Attwood, D.; Booth, C.<br />

Int. J. Pharm. 2005, 300, 22–31.<br />

(6) Booth, C.; Attwood, D. Macromol. Rapid Commun. 2000, 21, 501–527.<br />

103


7108 Langmuir, Vol. 24, No. 14, 2008 Juárez et al.<br />

formed by propylene oxide 6–9 <strong>and</strong> butylene oxide 10–13 with<br />

diblock <strong>and</strong> triblock architectures. On the other h<strong>and</strong>, formation<br />

<strong>of</strong> vesicular structures for poly(oxyalkylene) block copolymers<br />

has <strong>of</strong>ten needed energy supply via mechanical, sonochemical,<br />

or electrical methods. 14,15 Nevertheless, some works have reported<br />

the spontaneous formation <strong>of</strong> vesicles, denoting that these types<br />

<strong>of</strong> structure are likely to form because <strong>of</strong> relatively large volumeto-length<br />

ratios <strong>of</strong> the hydrophobic block <strong>and</strong> the relative lengths<br />

<strong>of</strong> the poly(oxyethylene) blocks. 16–18<br />

In the case <strong>of</strong> block copolymers containing poly(oxystyrene)<br />

blocks, the formation <strong>of</strong> elongated micelles has been unusual,<br />

whereas vesicle formation has not been reported. Probably, neither<br />

the composition nor the solvent conditions <strong>of</strong> the analyzed<br />

poly(ethylene oxide)-polystyrene oxide block copolymers promote<br />

the clear formation <strong>of</strong> other aggregate structures than<br />

spherical micelles. Only Yang et al. 19 have compared the<br />

association number (Nw) <strong>of</strong> copolymer micelles <strong>of</strong> copolymers<br />

E17B8 <strong>and</strong> E17S8 <strong>and</strong> found that substitution <strong>of</strong> the B block by<br />

the S block brought about a large rise in Nw, which was larger<br />

than the theoretical increase expected for spherical micelles.<br />

They hypothesized that the difference in Nw between these two<br />

copolymers might be explained in terms <strong>of</strong> micelle elongation,<br />

but no direct measurements were done. Moreover, in a recent<br />

report Elsabahy et al. 20 studied the drug solubilization capacity<br />

<strong>and</strong> release kinetics in vitro <strong>of</strong> the anticancer drug docetaxel by<br />

poly(oxystyrene)-poly(oxyethylene) block copolymers. These<br />

authors have demonstrated the existence <strong>of</strong> an elongated structure<br />

<strong>of</strong> micelles formed by copolymers E45S15 <strong>and</strong> E45S26 by TEM.<br />

Nevertheless, these authors also pointed out the possible influence<br />

<strong>of</strong> residual styrene oxide homopolymer from the synthesis step<br />

dissolved in the micelle core to elongate the aggregate structure,<br />

as discussed earlier by Nagarajan. 21<br />

Thus, in the present work, we studied the self-assembly<br />

properties <strong>and</strong> the different resulting aggregate structures <strong>of</strong> three<br />

copolymers with poly(oxyethylene) <strong>and</strong> poly(oxystyrene) blocks<br />

(E12S10,E10S10E10, <strong>and</strong> E137S18E137), one diblock <strong>and</strong> two triblock<br />

copolymers, respectively, with very different volume-to-length<br />

ratios <strong>of</strong> the hydrophobic block <strong>and</strong> relative lengths <strong>of</strong> the<br />

poly(oxyethylene) blocks, which should help in controlling<br />

aggregate morphology. The study was conducted by using<br />

different techniques such as surface tension, light scattering,<br />

transmission <strong>and</strong> scanning electron microscopies, polarized optical<br />

microscopy, <strong>and</strong> rheology. In fact, for copolymer E12S10<br />

(7) Chu, B.; Zhou, Z. Physical Chemistry <strong>of</strong> Polyoxyalkylene Block Copolymer<br />

Surfactants. In Nonionic Surfactants: Polyoxyalkylene Block Copolymers; Nace,<br />

V. M., Ed.; Surfactant Science Series 60; Marcel Dekker: New York, 1996.<br />

(8) Hvidt, S.; Jørgensen, E. B.; Brown, W.; Schillén, K. J. Phys. Chem. 1994,<br />

98, 12320–12328.<br />

(9) Gangaly, R.; Awal, V. K.; Hassan, P. A. J. Colloid Interface Sci. 2007,<br />

325, 693–700.<br />

(10) Kelarakis, A.; Havredaki, V.; Booth, C.; Nace, V. M. Macromolecules<br />

2002, 35, 5591–5594.<br />

(11) Chaibundit, C.; Ricardo, N. M. P. S.; Crothers, M.; Booth, C. Langmuir<br />

2002, 18, 4277–4283.<br />

(12) Li, H.; Yu, G.-E.; Price, C.; Booth, C.; Fairclough, J. P. A.; Ryan, A. J.;<br />

Mortensen, K. Langmuir 2003, 19, 1075–1081.<br />

(13) Chaibundit, C.; Sumnatrakool, P.; Chinchew, S.; Kanatharana, P.;<br />

Tattershall, C. E.; Booth, C.; Yuan, X.-F. J. Colloid Interface Sci. 2005, 283,<br />

544–554.<br />

(14) Battaglia, G.; Ryan, A. J. J. Am. Chem. Soc. 2005, 127, 8757–8764.<br />

(15) Zipfel, J.; Lindner, M.; Tsinou, M.; Alex<strong>and</strong>ridis, P.; Richtering, W.<br />

Langmuir 1999, 15, 2599–2602.<br />

(16) Battaglia, G.; Ryan, A. J. Nature Mat. 2005, 4, 869–876.<br />

(17) Bryskhe, K.; Jansson, J.; Topgaard, D.; Schillén, K.; Olsson, U. J. Phys.<br />

Chem. B 2004, 108, 9710–9719.<br />

(18) Harris, J. K.; Rose, G. D.; Bruening, M. L. Langmuir 2002, 18, 5337–<br />

5342.<br />

(19) Yang, Z.; Crothers, M.; Attwood, D.; Collet, J. H.; Ricardo, N. P. S.;<br />

Martini, L. G. A.; Booth, C. J. Colloid Interface Sci. 2003, 263, 312–317.<br />

(20) Elsabahy, M.; Perron, M.-E.; Bertr<strong>and</strong>, N.; Yu, G.-E.; Leroux, J.-C.<br />

Biomacromolecules 2007, 8, 2250–2257.<br />

(21) Nagarajan, R. Colloids Surf., B 1999, 16, 55–72.<br />

Table 1. Molecular Characteristics <strong>of</strong> the Copolymer a<br />

Mn/g mol -1<br />

(NMR)<br />

wt%S<br />

(NMR)<br />

Mw/Mn<br />

(GPC) Mw/g mol -1<br />

coexistence <strong>of</strong> spherical micelles with vesicular structures is<br />

detected. To the best <strong>of</strong> our knowledge, the spontaneous formation<br />

<strong>of</strong> vesicles by a poly(oxystyrene)-poly(oxyethylene) block<br />

copolymers is reported for the first time. On the other h<strong>and</strong>,<br />

formation <strong>of</strong> elongated micelles upon self-assembly <strong>of</strong> copolymer<br />

E10S10E10 in dilute solution is elucidated from light scattering<br />

<strong>and</strong> transmission electron techniques. In the case <strong>of</strong> copolymer<br />

E137S18E137, typical spherical micelles are observed as expected.<br />

The micellization properties <strong>of</strong> the latter copolymer had been<br />

already partially studied previously, <strong>and</strong> the present study<br />

completes the gap. 22,23 Finally, comparison <strong>of</strong> the self-assembly<br />

properties <strong>of</strong> the present copolymers with structurally related<br />

poly(ethylene oxide)-propylene oxide <strong>and</strong> poly(ethylene oxide)butylene<br />

oxide block copolymers is also made.<br />

Experimental Section<br />

Materials. The synthesis <strong>of</strong> the copolymers was described<br />

previously in detail. 1,2 Table 1 shows the molecular characteristics<br />

<strong>of</strong> the copolymers. Water was double distilled <strong>and</strong> degassed before<br />

use.<br />

Clouding. Solutions in small tubes (10 mm in diameter) were<br />

prepared by weighting copolymer <strong>and</strong> mixing with water in the<br />

mobile state before being stored at low temperature for several days<br />

(T ≈ 5 °C). Tubes containing 0.5 cm3 <strong>of</strong> solution were immersed<br />

in a water bath, which was heated at a heating/cooling rate <strong>of</strong> 0.2<br />

°C min-1 E12S10 1660 73 1.05 1760<br />

E10S10E10 1980 55 1.06 2130<br />

E137S18E137 14200 15 1.06 15100<br />

a<br />

Estimated uncertainty: Mn to (3%;wt%Sto(1%, Mw/Mn to (0.01.<br />

Mw calculated from Mn <strong>and</strong> Mw/Mn.<br />

from5to90°C. Clouding was detected by the eye. The<br />

tubes were also inverted to check the possibility <strong>of</strong> gel formation.<br />

The change from a mobile to an immobile system (or vice versa)<br />

was determined by inverting the tube at 1-min intervals.<br />

Surface Tension Measurements. Surface tensions (γ) <strong>of</strong>the<br />

copolymers were measured by the Wilhelmy plate method using<br />

Krüss K-12 surface tension equipment, equipped with a processor<br />

to acquire the data automatically. The equipment was connected to<br />

a circulating water bath to keep the temperature constant at the<br />

corresponding temperature to within (0.01 °C. The plate was cleaned<br />

by being washed with doubly distilled water followed by being<br />

heated in an alcohol flame. A copolymer stock solution was prepared<br />

with distilled water <strong>and</strong> diluted as required. For measurements, a<br />

solution was equilibrated at the proper temperature <strong>and</strong> the surface<br />

tension was recorded at 15-min intervals until a constant value was<br />

reached, a process that took 12-36 h depending on concentration.<br />

The accuracy <strong>of</strong> measurement was checked by frequent determination<br />

<strong>of</strong> the surface tension <strong>of</strong> pure water.<br />

Light Scattering Measurements. Dynamic <strong>and</strong> static light<br />

scattering (DLS <strong>and</strong> SLS) intensities were measured for copolymer<br />

solutions at 20 <strong>and</strong> 50 °C for E137S18E137 by means <strong>of</strong> an ALV-<br />

5000F (ALV-GmbH) instrument with vertically polarized incident<br />

light <strong>of</strong> wavelength λ ) 532 nm supplied by a CW diode-pumped<br />

Nd; YAG solid-state laser supplied by Coherent, Inc. <strong>and</strong> operated<br />

at 400 mW. The intensity scale was calibrated against scattering<br />

from toluene. Measurements were made at the scattering angle, θ,<br />

<strong>of</strong> 90° to the incident beam. Solutions were equilibrated at each<br />

chosen temperature for 30 min before we made a measurement.<br />

Experiment duration was in the range 5-15 min, <strong>and</strong> each experiment<br />

was repeated two or more times. To eliminate dust, samples were<br />

filtered through Millipore Millex filters (Triton free, 0.45-0.8-µm<br />

porosity).<br />

(22) Ricardo, N. M. P. S.; Chaibundit, C.; Yang, Z.; Attwood, D.; Booth, C.<br />

Langmuir 2006, 22, 1301–1306.<br />

(23) Pinho, M. E. N.; Costa, F. M. L. L.; Filho, F. B. S.; Ricardo, N. M. P. S.;<br />

Yeates, S. G.; Attwood, D.; Booth, C. Int. J. Pharm. 2007, 328, 95–98.<br />

104


<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> OCH2CH(C6H5)-OCH2CH2 Block Copolymers Langmuir, Vol. 24, No. 14, 2008 7109<br />

The correlation functions from DLS were analyzed by the CONTIN<br />

method to obtain intensity distributions <strong>of</strong> decay rates (Γ). 24 The<br />

decay rate distributions gave distributions <strong>of</strong> apparent diffusion<br />

coefficient (Dapp ) Γ/q 2 , q ) (4πns/λ) sin(θ/2), where ns ) refractive<br />

index <strong>of</strong> solvent), <strong>and</strong> integrating over the intensity distribution<br />

gave the intensity-weighted average <strong>of</strong> Dapp. Values <strong>of</strong> the apparent<br />

hydrodynamic radius (rh,app, radius <strong>of</strong> hydrodynamically equivalent<br />

hard sphere corresponding to Dapp) were calculated from the<br />

Stokes-Einstein equation<br />

r h,app ) kT/(6πηD app)<br />

where k is the Boltzmann constant <strong>and</strong> η is the viscosity <strong>of</strong> water<br />

at temperature T.<br />

The basis for analysis <strong>of</strong> SLS was the Debye equation<br />

mic<br />

K * c/(I - Is ) ) 1⁄Mw + 2A2c + ... (2)<br />

where I is intensity <strong>of</strong> light scattering from solution relative to that<br />

from toluene, Is is the corresponding quantity for the solvent, c is<br />

the concentration (in g dm-3 ), Mw mic is the mass-average molar mass<br />

<strong>of</strong> the solute, A2 is the second virial coefficient (higher coefficients<br />

being neglected), <strong>and</strong> K* is the appropriate optical constant which<br />

includes the specific refractive index increment, dn/dc. Values <strong>of</strong><br />

the specific refractive index increment, dn/dc, <strong>and</strong> its temperature<br />

increment were taken from previous reports. 1,2 The possible effect<br />

<strong>of</strong> the different refractive indices <strong>of</strong> the blocks on the derived molar<br />

masses <strong>of</strong> micelles has been found to be negligible. 25 Other quantities<br />

used were the Raleigh ratio <strong>of</strong> toluene for vertically polarized light,<br />

Rv ) 2.57 × 10-5 [1 + 3.68 × 10-3 (t - 25)] cm-1 (t in °C) <strong>and</strong> the<br />

refractive index <strong>of</strong> toluene, n ) 1.4969[1-5.7 × 10-4 (t - 20)]. 26<br />

Polarized Optical Microscopy (POM). POM experiments were<br />

made by using a Leica TCS-SP2 microscope equipped with cross<br />

polarizers in copolymer sample before being filtered. Samples after<br />

filtration did not allow measurements <strong>of</strong> the typical lamellae patterns<br />

because <strong>of</strong> insufficient instrument resolution.<br />

Transmission Electron Microscopy (TEM). Samples for TEM<br />

were prepared by evaporation <strong>of</strong> a 2.5 g dm-3 aqueous micellized<br />

filtered copolymer solutions under air (with 0.8-µm pore filters). A<br />

drop <strong>of</strong> copolymer solution was placed on an electron microscope<br />

copper grid <strong>and</strong> stained with 2% (v/v) <strong>of</strong> phosphotungstic acid.<br />

After drying, electron micrographs <strong>of</strong> the sample were recorded<br />

with a Phillips CM-12 electron microscope.<br />

Cryo-Scanning Electron Microscopy (Cryo-SEM). Cryo-<br />

SEM micrographs were acquired by using a JEOL JSM-6360LV<br />

microscope with accelerating voltage <strong>of</strong> 0.5-30 keV. Solutions were<br />

filtered <strong>and</strong> flash frozen in liquid ethane <strong>and</strong> stored in liquid nitrogen.<br />

They were transferred to a cryo preparation chamber at -90 °C <strong>and</strong><br />

held at this temperature for 2 min in a vacuum to sublime the surface<br />

ice <strong>and</strong> sputter-coated with conductive chromium.<br />

Rheometry. The rheological properties <strong>of</strong> the samples were<br />

determined using a Bohlin CS10 rheometer with water bath<br />

temperature control. Couette geometry (bob, 24.5 mm in diameter,<br />

27 mm in height; cup, 26.5 mm in diameter, 29 mm in height) was<br />

used, with 2.5 cm3 <strong>of</strong> sample being added to the cup in the mobile<br />

state. Samples <strong>of</strong> very high modulus were investigated using cone<strong>and</strong>-plate<br />

geometry (diameter 40 mm, angle 4°). A solvent trap<br />

maintained a water-saturated atmosphere around the cells. Frequency<br />

scans <strong>of</strong> storage (G′) <strong>and</strong> loss (G″) moduli were recorded for selected<br />

copolymer concentrations <strong>and</strong> temperatures with the instrument in<br />

oscillatory shear mode <strong>and</strong> with the strain amplitude (A) maintained<br />

at a low value (A < 0.5%) by means <strong>of</strong> autostress facility <strong>of</strong> the<br />

(24) Provencher, S. W. Makromol. Chem. 1979, 180, 201–209.<br />

(25) Mai, S.-M.; Booth, C.; Kelarakis, A.; Havredaki, V.; Ryan, J. A. Langmuir<br />

2000, 16, 1681–1688.<br />

(26) (a) El-Kashef, H. ReV. Sci. Instrum. 1998, 69, 1243–1245. (b) Liu, T.;<br />

Schuch, H.; Gerst, M.; Chu, B. Macromolecules 1999, 32, 6031–6042. (c) The<br />

temperature dependence <strong>of</strong> the Raleigh ratio was not found in the literature <strong>and</strong>,<br />

as a working approximation, values <strong>of</strong> Rv for toluene were adjusted relative to<br />

those published for benzene. Gulari, E.; Chu, B.; Liu, T. Y. Biopolymers 1979,<br />

18, 2943. The adjustment is small over the temperature range involved.<br />

(1)<br />

Figure 1. Temperatures at which turbidity was visually observed on<br />

heating dilute solutions <strong>of</strong> copolymers (9)E12S10 <strong>and</strong> (O)E10S10E10. The<br />

lines are to guide the eye.<br />

Table 2. Critical Micelle Concentration (cmc), Surface Excess<br />

Concentration (Γmax), Minimum Area per Molecule (Amin), Gibbs<br />

°<br />

Free Energy <strong>of</strong> Micelle Formation (∆Gmic),<br />

<strong>and</strong> Gibbs Free<br />

°<br />

Energy <strong>of</strong> Adsorption (∆Gads)<br />

<strong>of</strong> the Present Copolymers a<br />

polymer T (°C) cmc (g dm -3 ) Γmax (10 -6 mol m -2 ) Amin (nm 2 )<br />

E12S10 15 0.006 0.97 1.7<br />

E10S10E10 15 0.012 0.91 1.8<br />

E137S18E137 15 0.017 0.40 4.1<br />

50 0.013 0.46 3.6<br />

a<br />

Estimated uncertainties: cmc, Γmax, Amin to (10%.<br />

Bohlin s<strong>of</strong>tware. This ensured that measurements <strong>of</strong> G′ <strong>and</strong> G″ were<br />

in the linear viscoelastic region. Temperature scans were recorded<br />

for selected copolymer concentrations at a frequency <strong>of</strong> 1 Hz. The<br />

samples were heated at ca. 0.5 °C in the range 5-80 °C.<br />

Measurements on solutions <strong>of</strong> low modulus (G″ ) 1-10 Pa) which<br />

felt outside the range for satisfactory autostress feedback were<br />

rejected.<br />

Results<br />

Clouding. Clouding temperatures obtained visually over a<br />

wide concentration range are plotted in Figure 1. The overall<br />

shape for copolymers E12S10 <strong>and</strong> E10S10E10 is as expected for<br />

solutes with a lower critical solution temperature. Solutions <strong>of</strong><br />

copolymer E12S10 were only transparent at very low concentration,<br />

<strong>and</strong> opacity increased in a narrow interval as the concentration<br />

increased. Copolymer E10S10E10 solutions were clear in a wider<br />

temperature <strong>and</strong> concentration range. Meanwhile, solutions <strong>of</strong><br />

copolymer E137S18E137 remained clear to the eye in the whole<br />

temperature <strong>and</strong> concentration range analyzed. As described in<br />

subsequent sections, the solutions were formed by copolymer<br />

aggregates over the concentration range involved <strong>and</strong> the detailed<br />

shape <strong>of</strong> the cloud point curve for the former copolymers has<br />

to be discussed in that context. This reflects the different behavior<br />

<strong>of</strong> the three copolymers as will be commented in detail below.<br />

We return to this point after presentation <strong>of</strong> the light scattering<br />

results. On the other h<strong>and</strong>, gelation was only observed for<br />

copolymer E137S18E137 at concentrations larger than 15% (w/v)<br />

with no lower temperature boundary.<br />

Critical Micelle Concentration <strong>and</strong> Surface Properties <strong>of</strong><br />

Copolymers. Surface tensions <strong>of</strong> solutions <strong>of</strong> samples corresponding<br />

to copolymers E137S18E137,E10S10E10, <strong>and</strong> E12S10 were<br />

investigated, as set out in Table 2. Plots <strong>of</strong> surface tension against<br />

log(concentration) are plotted in Figure 2. The concentration at<br />

105


7110 Langmuir, Vol. 24, No. 14, 2008 Juárez et al.<br />

Figure 2. Surface tension (γ) vs logarithm <strong>of</strong> concentration (g dm -3 )<br />

<strong>of</strong> copolymers (O) E10S10E10 <strong>and</strong> (9) E12S10 at 15 °C <strong>and</strong> E137S18E137 at<br />

(2) 15 <strong>and</strong> (3) 50°C in aqueous solution.<br />

which the surface tension reached a steady value served to define<br />

the cmc. Representative values <strong>of</strong> the cmc <strong>and</strong> <strong>of</strong> the surface<br />

tension at the cmc (γcmc) are listed in Table 2. It can be observed<br />

that the triblock copolymers have higher cmc values than the<br />

diblock copolymer. The effect is originated by the entropy <strong>of</strong><br />

the triblock chains constrained by two block junctions in the<br />

core-shell interface <strong>of</strong> the aggregate, compared with only one<br />

constrain for the diblock chain. The present values for solutions<br />

<strong>of</strong> copolymers E137S18E137, E10S10E10, <strong>and</strong> E12S10 at 15 °C are<br />

in good agreement with previously reported values for other<br />

diblock <strong>and</strong> triblock copolymers <strong>of</strong> ethylene oxide <strong>and</strong> styrene<br />

oxide. 27,28<br />

On the other h<strong>and</strong>, the surface excess concentration, Γmax, <strong>and</strong><br />

the minimum area per copolymer molecule, Αmin, at the air/<br />

water interface were obtained using the surface tension measurements<br />

<strong>and</strong> the following equations:<br />

Γmax ) - 1<br />

RT( ∂γ<br />

∂ ln c) T,p<br />

Amin ) 1/NAΓmax where R is the gas constant, NA is Avogadro’s number, <strong>and</strong> c is<br />

the copolymer concentration.<br />

Because <strong>of</strong> their higher hydrophobicity, copolymers E12S10 <strong>and</strong><br />

E10S10E10 display a larger interface coverage than E137S18E137, whose<br />

molecules also need more space to be accommodated because <strong>of</strong><br />

the longer length. 28 On the other h<strong>and</strong>, for the triblock E137S18E137<br />

there is a decrease in the value <strong>of</strong> A as the temperature rises as a<br />

consequence <strong>of</strong> the increase in the thermal motions <strong>of</strong> the polymer<br />

chains <strong>and</strong> the dehydration <strong>of</strong> the hydrophilic blocks. 29,30<br />

(27) Booth, C.; Attwood, D.; Price, C. Phys. Chem. Chem. Phys. 2006, 8,<br />

3612–3622.<br />

(28) Barbosa, S.; Cheema, M. A.; Taboada, P.; Mosquera, V. J. Phys. Chem.<br />

B 2007, 111, 10920–10928.<br />

(29) Kelarakis, A.; Havredaki, V.; Yu, G.-E.; Derici, L.; Booth, C. Macromolecules<br />

1998, 31, 944–946.<br />

(3)<br />

(4)<br />

Figure 3. Intensity fraction distributions <strong>of</strong> the apparent hydrodynamic<br />

radius for (a) copolymer E137S18E137 (60 g/dm -3 at 20 °C) <strong>and</strong> (b)<br />

copolymer E12S10 (2.5 g/dm -3 at 10 °C).<br />

Vesicular Formation in Copolymer E12S10. Intensity fraction<br />

distributions <strong>of</strong> decay times for the three copolymers are<br />

shown in Figure 3. In the case <strong>of</strong> copolymers E10S10E10 <strong>and</strong><br />

E137S18E137, the distribution contained one single peak (see Figure<br />

3a for copolymer E137S18E137 as an example). In contrast, three<br />

single narrow peaks were observed for copolymer E12S10 at<br />

suitable concentrations to be measured by DLS (Figure 3b). These<br />

peaks did not vary in a systematic way in the narrow copolymer<br />

concentration range analyzed (0.5-3 gdm -3 ) provided that<br />

solutions were not completely clear (turbidity increased as the<br />

copolymer concentration rises), so a true hydrodynamic radius<br />

could not be obtained (plot not shown). Because <strong>of</strong> the translucent<br />

aspect <strong>of</strong> these solutions <strong>and</strong> to know if these solutions were<br />

isotropic or not, we analyzed some <strong>of</strong> the samples by POM.<br />

Surprisingly, the typical patterns displayed by spherical lamellae<br />

were observed for all the samples in the concentration range<br />

anlyzed 31 (Figure 4a).<br />

In contrast, no presence <strong>of</strong> these patterns was observed for<br />

solutions <strong>of</strong> copolymer E10S10E10 or copolymer E137S18E137 in<br />

the concentration range analyzed, even in solutions not optically<br />

clear. The different sizes <strong>of</strong> the observed circular patterns in<br />

polarized spectroscopy denote the polydisperse size <strong>of</strong> the vesicles.<br />

It is necessary to note that samples for POM experiments were<br />

unfiltered samples provided that filtered ones did not allow<br />

accurate measurements because <strong>of</strong> instrument resolution. Thus,<br />

sizes <strong>of</strong> vesicles from POM are larger than those derived from<br />

electron microscopies (between 1 <strong>and</strong> 5 µm) <strong>and</strong> even more<br />

polydisperse. The presence <strong>of</strong> vesicular structures was further<br />

corroborated by TEM <strong>and</strong> cryo-SEM, as seen in Figure 4b,c, in<br />

this case using filtered solutions as for DLS. Despite the inherent<br />

problems <strong>of</strong> TEM due to solvent evaporation during sample<br />

preparation, which affects sizes <strong>and</strong> structures <strong>of</strong> nanostructures<br />

in solution, the coexistence <strong>of</strong> micelles with a mean size <strong>of</strong> ca.<br />

(30) Sony, S. S.; Sastry, N. V.; George, J.; Bohidar, H. B. Langmuir 2003,<br />

19, 4597–4603.<br />

(31) Kunieda, H.; Nakamura, K.; Davis, H. T.; Evans, D. F. Langmuir 1991,<br />

7, 1915–1921.<br />

106


<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> OCH2CH(C6H5)-OCH2CH2 Block Copolymers Langmuir, Vol. 24, No. 14, 2008 7111<br />

Figure 4. (a) POM image <strong>of</strong> vesicles formed by copolymer E12S10. Picture was taken at 100× magnification. (b) TEM images <strong>of</strong> vesicles (denoted<br />

with arrows) that coexist with smaller micelles. (c) Cryo-SEM picture <strong>of</strong> polymer vesicles <strong>of</strong> copolymer E12S10. The smallest vesicles are denoted<br />

with arrows.<br />

Figure 5. Concentration dependence <strong>of</strong> the reciprocal apparent<br />

hydrodynamic radii for micellar solutions <strong>of</strong> copolymers E137S18E137 at<br />

(9) 15 <strong>and</strong> (O) 50°C <strong>and</strong> copolymer E10S10E10 at 15 °C (inset).<br />

16 nm <strong>and</strong> some small vesicles <strong>of</strong> polydisperse size (mean size<br />

<strong>of</strong> ca. 65 nm; see arrows in Figure 4b) can be discerned. Cryo-<br />

SEM showed also the existence <strong>of</strong> larger vesicles (with sizes<br />

around ca. 400 nm) at more concentrated copolymer solutions,<br />

as seen in Figure 4c, which coexist with smaller vesicles <strong>of</strong> ca.<br />

100 nm <strong>of</strong> diameter (see arrows). In this figure, pores created<br />

from vesicle disintegration during sample preparation can be<br />

also observed. Micelles were not seen because the equipment<br />

Figure 6. Debye plots for micellar solutions <strong>of</strong> (a) copolymer E137S18E137<br />

at (9) 15 <strong>and</strong> (O) 50°C <strong>and</strong> (b) copolymer E10S10E10 at 15 °C.<br />

resolution was not sufficient. Then, comparing the size distributions<br />

derived from DLS with the sizes <strong>of</strong> the structures seen by<br />

TEM <strong>and</strong> SEM, we can assign the first peak <strong>of</strong> the DLS size<br />

107


7112 Langmuir, Vol. 24, No. 14, 2008 Juárez et al.<br />

Table 3. Micellar <strong>and</strong> Hydration Properties <strong>of</strong> Block Copolymers E12S10, E10S10E10, <strong>and</strong> E137S18E137<br />

polymer T/°C A2/10 -4 mol g -2 cm 3 δt Mw mic /10 5 mol g -1 rh/nm Nw rt/nm<br />

E10S10E10 15 4.5 1.0 4.7 20.2 202 21.5<br />

E137S18E137 15 0.2 5.8 4.8 12.9 34 9.9<br />

50 0.27 4.8 6.3 11.8 42 10.2<br />

distribution <strong>of</strong> copolymer E12S10 to micelles provided that this<br />

species lies in the typical sizes <strong>of</strong> copolymer micelles (with an<br />

apparent radius <strong>of</strong> ca. 13 nm), whereas the other two peaks are<br />

assigned to polymer vesicles <strong>of</strong> different sizes. In this regard,<br />

differences in vesicle sizes between SEM, DLS, <strong>and</strong> TEM may<br />

arise from solvent evaporation in the latter technique, which<br />

dehydrates the polymer corona <strong>and</strong> shrinks the polymer structure.<br />

On the other h<strong>and</strong>, polymer solutions when measured by TEM<br />

showed the presence <strong>of</strong> polymer aggregated material, which may<br />

be composed <strong>of</strong> the largest vesicles as a consequence <strong>of</strong> the<br />

sample evaporation process since these were not observed by<br />

this technique (picture not shown).<br />

Micellar Properties <strong>of</strong> Copolymers E137S18E137 <strong>and</strong><br />

E10S10E10. Hydrodynamic Radii. Now we will focus on the<br />

micellar properties <strong>of</strong> copolymers E137S18E137 <strong>and</strong> E10S10E10. Plots<br />

<strong>of</strong> the reciprocal <strong>of</strong> the intensity average <strong>of</strong> rh,app (by using the<br />

Stokes-Einstein relation) are shown in Figure 5 for copolymers<br />

E137S18E137 <strong>and</strong> E10S10E10. The intercepts <strong>of</strong> these plots at c )<br />

0 gave values <strong>of</strong> rh. Note that through eq 1 the reciprocal <strong>of</strong> rh,app<br />

is proportional to Dapp/ηT <strong>and</strong> thus compensated for changes in<br />

solvent viscosity <strong>and</strong> temperature. The positive slopes in this<br />

figure are consistent with these micelles acting effectively as<br />

hard spheres. 1 This behavior is usually accommodated by<br />

introducing a diffusion second virial coefficient (kd) in the equation<br />

<strong>of</strong> the straight line<br />

D app ) D(1 + k d c + ...) (5)<br />

The coefficient kd is related to the thermodynamic second<br />

virial coefficient A2 by 32<br />

mic<br />

kd ) 2A2Mw - kf - 2V (6)<br />

where kf is the friction coefficient <strong>and</strong> V is the specific volume<br />

<strong>of</strong> the micelles in solution. As seen from Figure 6, the positive<br />

term in eq 7 is dominant. Relating A2 to the effective hard-sphere<br />

volume <strong>of</strong> the micelles, Vhs (A2 ) 4NAVhs/Mw 2 (mic) ) 33 shows that<br />

the first term depends on the ratio Vhs/Mw mic . A2 values (Table<br />

3) are positive in agreement with the commented micellar<br />

behavior. On the other h<strong>and</strong>, the larger insensitivity <strong>of</strong> 1/rh,app<br />

with changes in concentration <strong>of</strong> copolymer E10S10E10 is<br />

associated to their higher molar masses <strong>and</strong> is a common feature<br />

observed for elongated micelles. 4,14<br />

Aggregation Numbers. As noted in the , the Debye equation<br />

(eq 2) is for nonideal dilute solution <strong>of</strong> particles that are small<br />

relative to the wavelength <strong>of</strong> the light. The Debye equation taken<br />

to the second term, A2, could not be used to analyze the SLS data<br />

for copolymer E137S18E137 as micellar interaction causes curvature<br />

<strong>of</strong> the Debye plot across the concentration range investigated<br />

(Figure 6). The fitting procedure used for these curves was based<br />

on scattering theory for hard spheres 34 whereby the interparticle<br />

interference factor (structure factor, S) in the scattering equation<br />

mic<br />

K * c ⁄(I - Is ) ) 1⁄SMw was approximated by<br />

1⁄S ) [(1 + 2φ) 2 - φ 2 (4φ - φ 2 )](1 - φ) -4<br />

(8)<br />

where φ is the volume fraction <strong>of</strong> equivalent uniform spheres.<br />

Values <strong>of</strong> φ were conveniently calculated from the volume fraction<br />

<strong>of</strong> copolymer in the system by applying a thermodynamc<br />

(7)<br />

expansion factor δt )Vt/Va, where Vt is the thermodynamic volume<br />

<strong>of</strong> a micelle (i.e., one-eighth <strong>of</strong> the volume, u, excluded by<br />

one micelle to another) <strong>and</strong> Va is the anhydrous volume <strong>of</strong> a<br />

micelle (Va )VMw mic /NA), where V is the partial specific volume<br />

<strong>of</strong> the copolymer solute. The fitting parameter, δt, applies as an<br />

effective parameter for compact micelles irrespective <strong>of</strong> their<br />

exact structure. The method is equivalent to using the virial<br />

expansion for the structure factor <strong>of</strong> effective hard spheres taken<br />

to its seventh term 34 but requires just two adjustable parameters<br />

(i.e., Mw mic <strong>and</strong> δt). Values <strong>of</strong> these two quantities are shown in<br />

Table 3. In addition, weight-average association numbers, Nw,<br />

<strong>of</strong> the copolymer micelles were calculated from the micellar<br />

molecular mass derived from the fitting procedure shown above<br />

<strong>and</strong> the weight-average molecular weight <strong>of</strong> the copolymer in<br />

the unimer state (i.e., Nw ) Mw(micelle)/Mw(molecule); Table<br />

1). Values <strong>of</strong> the equivalent hard-sphere radius (the thermodynamic<br />

radius, rt) calculated from the thermodynamic volume <strong>of</strong><br />

the micelles (i.e., from Vt ) δtVa) are also listed in Table 3. It<br />

is also clear that the effective hard-sphere radius, like the<br />

hydrodynamic radius, has a more direct relation to spherical<br />

micelles than to other shapes.<br />

As expected, Nw values for copolymer E137S18E137 slightly<br />

increase as the temperature rises since water becomes a poorer<br />

solvent for the poly(oxyethylene) blocks. This is consistent with<br />

the effect <strong>of</strong> temperature in micelles <strong>of</strong> other copoly(oxyalkylene)s.<br />

28,35 In addition, present aggregation numbers are in<br />

agreement with previously reported ones for this copolymer at<br />

other temperatures. 23<br />

TEM imaging was undertaken to directly visualize the micelle<br />

shape. Copolymer E137S18E137 formed nearly spherical micelles<br />

in solution (Figure 7a), whereas copolymer E10S10E10 selfassembled<br />

into spheroidal aggregates (Figure 7b). Under TEM,<br />

the diameters <strong>of</strong> micelles were close to those obtained by DLS.<br />

The mean length <strong>of</strong> the major axis <strong>of</strong> the spheroids was ca. 22<br />

nm. Nevertheless, one has to bear in mind that with TEM we<br />

image single particles whereas DLS gives an average size<br />

estimation, which is biased toward the larger-size end <strong>of</strong> the<br />

population distribution. In addition, it should be pointed out that<br />

rh for elongated micelles derived from DLS is an effective value<br />

for “equivalent spheres”, <strong>and</strong> hence rh is not expected to equal<br />

the ellipsoid or cylinder radius.<br />

Rheological Behavior. Solutions <strong>of</strong> copolymers E10S10E10<br />

<strong>and</strong> E137S18E137 were analyzed to study their rheological behavior.<br />

Solutions <strong>of</strong> E10S10E10 (0-50% (w/v)) were investigated for gel<br />

formation using tube inversion experiments. Immobility implies<br />

no detectable flow over a period <strong>of</strong> hours or days. To a good<br />

approximation, immobility in the test requires the gel to have a<br />

yield stress higher than 30 Pa. 36 Adopting the notation used by<br />

Hvidt et al., 36–38 we refer to this immobile phase as “hard gel”.<br />

(32) Vink, H. J. Chem. Soc., Faraday Trans. 1 1985, 81, 1725–1730.<br />

(33) Flory, P. J. Principles <strong>of</strong> Polymer Chemistry; Cornell University Press:<br />

Ithaca, NY, 1953.<br />

(34) (a) Percus, J. K.; Yevick, G. J. J. Phys. ReV. 1958, 110, 1–13. (b) Vrij,<br />

A. J. Chem. Phys. 1978, 69, 1742–1747. (c) Carnahan, N. F.; Starling, K. E.<br />

J. Chem. Phys. 1969, 51, 635–636.<br />

(35) Booth, C.; Attwood, D. Macromol. Rapid Commun. 2000, 21, 501–527.<br />

(36) Hvidt, S.; Jørgensen, E. B.; Brown, W.; Schillén, K. J. Phys. Chem. 1994,<br />

98, 12320–12328.<br />

(37) Kelarakis, A.; Mingvanish, W.; Daniel, C.; Li, H.; Havredaki, V.; Heatley,<br />

F.; Booth, C.; Hamley, I. W. Phys. Chem. Chem. Phys. 2000, 2, 2755–2763.<br />

108


<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> OCH2CH(C6H5)-OCH2CH2 Block Copolymers Langmuir, Vol. 24, No. 14, 2008 7113<br />

Figure 7. TEM micrograhs <strong>of</strong> micelles formed by copolymers (a) E137S18E137 <strong>and</strong> (b) E10S10E10.<br />

Figure 8. Aqueous solutions <strong>of</strong> copolymer E137S18E137. (a) Temperature<br />

dependence <strong>of</strong> (9) storage (G′) <strong>and</strong> (O) loss (G″) moduli for aqueous<br />

hard gel at 50% (w/v), frequency, f ) 1 Hz, <strong>and</strong> strain amplitude, A )<br />

0.5%. (b) Temperature dependence <strong>of</strong> storage (G′) (9), <strong>and</strong> loss (G″),<br />

(O), moduli at 10% (w/v), f ) 1 Hz, <strong>and</strong> A ) 0.5%.<br />

They have used yield stress (σy) <strong>and</strong> modulus (storage <strong>and</strong> loss,<br />

G′ <strong>and</strong> G″) to define three convenient subdivisions <strong>of</strong> the range<br />

<strong>of</strong> behaviors observed: immobile hard gel (high σy <strong>and</strong> G′, G′<br />

> G″), mobile s<strong>of</strong>t gel (low σy <strong>and</strong> G′, G′ > G″), <strong>and</strong> mobile<br />

sol (null σy <strong>and</strong> very low G′, G′ < G″).<br />

Temperature Dependence <strong>of</strong> Copolymer E137S18E137. The solhard<br />

gel boundary was already determined by tube inversion <strong>and</strong><br />

rheology previously, 39 but additional measurements were done<br />

to complete its phase diagram. Figure 8a shows an example <strong>of</strong><br />

a temperature scan <strong>of</strong> the logarithm <strong>of</strong> the elastic modulus for<br />

a concentrated solution above the critical gel concentration (cgc)<br />

for this copolymer (19% (w/v), derived from the expression cgc<br />

) 10 2 Faφc/δt, where φc ) 0.68 is the volume fraction <strong>of</strong> spherical<br />

micelles packed in a body-centered structure. 39 On the basis <strong>of</strong><br />

a critical value <strong>of</strong> G″ ) 1 kPa for a hard gel, the 50% (w/v)<br />

(38) Almgrem, M.; Brown, W.; Hvidt, S. Colloid Polym. Sci. 1995, 273, 2–15.<br />

(39) Hamley, I. W.; Castelletto, V.; Ricardo, N. P. M. S.; Pinho, M. E. N.;<br />

Booth, C.; Attwood, D.; Yang, Z. Polym. Int. 2007, 56, 88–92.<br />

Figure 9. Phase diagram <strong>of</strong> copolymer E137S18E137 from (0) tube inversion<br />

<strong>and</strong> (9, b) rheology showing the sol, s<strong>of</strong>t gel, <strong>and</strong> hard gel regions.<br />

copolymer solution is a hard gel in the whole range, with a<br />

maximum value <strong>of</strong> G′ ≈ 16 kPa, <strong>and</strong> G′ > G″ throughout the<br />

whole experiment. In Figure 8b, a temperature scan <strong>of</strong> G′ <strong>and</strong><br />

G″ at a concentration below the hard-gel boundary (10% w/v)<br />

is also shown as an example. Other polymer concentrations in<br />

this region below the cgc display the same pr<strong>of</strong>ile (not shown).<br />

Using the conditions described above, we found that at<br />

temperatures at which G′ (1 Hz) exceeds 10 Pa the solutions are<br />

s<strong>of</strong>t gels, otherwise sols. 36–38 As illustrated, within the region <strong>of</strong><br />

raised modulus G′ was higher than G″, which justifies the<br />

convenient term “s<strong>of</strong>t gel” for the fluid to distinguish it from sol.<br />

A complete phase diagram <strong>of</strong> the regions <strong>of</strong> sol, s<strong>of</strong>t gel, <strong>and</strong><br />

hard gel defined by rheometry <strong>and</strong> tube inversion for copolymer<br />

E137S18E137 is shown in Figure 9. These sol-gel diagrams follow<br />

the general pattern as those published previously for other EO/<br />

SO block copolymers. 1,2 The upper limit to the s<strong>of</strong>t gel region<br />

for the copolymer is not reached within the temperature range<br />

investigated, consistent with the stability at high temperature <strong>of</strong><br />

the hard gels <strong>of</strong> this copolymer.<br />

Frequency Dependence <strong>of</strong> Copolymer E137S18E137. To get a<br />

more detailed picture about the rheological behavior <strong>of</strong> both<br />

hard <strong>and</strong> s<strong>of</strong>t gels, frequency scans <strong>of</strong> solution <strong>of</strong> E137S18E137<br />

block copolymer were performed. Figure 10a shows the frequency<br />

scan obtained for the 15% (w/v) solution <strong>of</strong> copolymer E137S18E137<br />

at 25 °C. This type <strong>of</strong> s<strong>of</strong>t gel at temperatures <strong>and</strong> concentrations<br />

relatively near the hard-gel boundary can be assigned as defective<br />

versions <strong>of</strong> the cubic packed hard gels (i.e., small structured<br />

domains in an overall fluid matrix). These s<strong>of</strong>t gels are<br />

109


7114 Langmuir, Vol. 24, No. 14, 2008 Juárez et al.<br />

Figure 10. Frequency dependence <strong>of</strong> (b) storage <strong>and</strong> (O) loss moduli<br />

(log-log plot) for (a) 15% (w/v) aqueous solution <strong>of</strong> copolymer<br />

E137S18E137 in the s<strong>of</strong>t gel region <strong>and</strong> (b) 6% (w/v) aqueous solution <strong>of</strong><br />

copolymer E137S18E137 in the s<strong>of</strong>t gel region at T ) 60 °C.<br />

characterized by the almost constant value <strong>of</strong> log G′, the minimum<br />

in log G″, <strong>and</strong> both moduli do not show a crossover point in the<br />

measured frequency range, all <strong>of</strong> which are characteristic <strong>of</strong> hard<br />

gels formed by cubic packing <strong>of</strong> spherical micelles, 40 although<br />

the modulus values are low compared to those <strong>of</strong> the hard gel<br />

(G′ ≈ 70 Pa <strong>and</strong> G″ ≈ 5Paatf ) 1 Hz, A ) 0.5%). S<strong>of</strong>t gels<br />

<strong>of</strong> this type have been also found for other EmSnEm,EmSn, 28 <strong>and</strong><br />

EmGnEm 4 (where G ) phenylglycidyl unit) copolymers studied<br />

by rheometry, <strong>and</strong> some EmBn copolymers using SAXS 41 <strong>and</strong><br />

Pluronics using SANS. 42<br />

On the other h<strong>and</strong>, the scan obtained for a 6% (w/v) solution<br />

at 60 °C differs from that <strong>of</strong> Figure 10a (Figure 10b), having very<br />

low values <strong>of</strong> G′ (e.g., G′ ≈ 8Paatf ) 1 Hz, A ) 0.5%) <strong>and</strong><br />

with a moduli crossover at low frequency, corresponding to a<br />

Maxwell fluid, at most, showing localized cubic order. This must<br />

be a consequence <strong>of</strong> the weak attraction <strong>of</strong> spherical micelles in<br />

water at temperatures where it is a poor solvent for the micelles.<br />

The transition from sol to s<strong>of</strong>t gel may well occur when aggregates<br />

<strong>of</strong> spherical micelles reach the percolation threshold, yielding<br />

sufficient structure to cause the characteristic rheological effect.<br />

S<strong>of</strong>t gels <strong>of</strong> this type have been also identified in aqueous micellar<br />

solutions <strong>of</strong> a wide range <strong>of</strong> copolymers including other EO/SO<br />

copolymers, 1,2,28,40 as well as in EmPnEm (P ) propylene oxide<br />

unit) by Hvidt <strong>and</strong> co-workers who ascribed the effect in certain<br />

systems to the intervention <strong>of</strong> the sphere-to-rod micellar<br />

transition, 36,38 <strong>and</strong> by Mallamace <strong>and</strong> co-workers who identified<br />

mechanisms <strong>of</strong> percolation <strong>and</strong> packing (structural-arrest),<br />

depending on temperature <strong>and</strong> concentration. 43,44<br />

(40) Li, H.; Yu, G.-E.; Price, C.; Booth, C.; Fairclough, J. P. A.; Ryan, A. J.;<br />

Mortensen, K. Langmuir 2003, 19, 1075–1081.<br />

(41) Castelletto, V.; Caillet, C.; Fundin, J.; Hamley, I. W.; Yang, Z.; Kelarakis,<br />

A. J. Chem. Phys. 2002, 116, 10947–10958.<br />

(42) Prud’homme, R. K.; Wu, G.; Schneider, D. K. Langmuir 1996, 12, 4651–<br />

4659.<br />

(43) Lobry, L.; Micali, N.; Mallamace, M.; Liao, C.; Chen, S. H. Phys. ReV.<br />

E 1999, 60, 7076–7087.<br />

(44) Chen, S. H.; Chen, W. R.; Mallamace, F. Science 2003, 300, 619–622.<br />

(45) Antonietti, M.; Forster, S. AdV. Mater. 2003, 15, 1323–1333.<br />

Figure 11. (a) Temperature dependence <strong>of</strong> (b) storage <strong>and</strong> (O) loss<br />

moduli (log-log plot) for a 20% (w/v) aqueous solution <strong>of</strong> copolymer<br />

E10S10E10. (b) Frequency dependence <strong>of</strong> storage (close symbols) <strong>and</strong><br />

loss moduli (open symbols) (log-log plot) for (b) 20, (9) 30, <strong>and</strong> (2)<br />

40% (w/v) aqueous solution <strong>of</strong> copolymer E10S10E10 at 15 °C.<br />

Concentration <strong>and</strong> Temperature Dependence <strong>of</strong> Copolymer<br />

E10S10E10. For copolymer E10S10E10, gelation was not found in<br />

the concentration range analyzed. Concentrations in the range<br />

5-30% (w/v) <strong>and</strong> temperatures in the range 10-45 °C were<br />

investigated despite the turbid solution, a signal that the phaseseparation<br />

boundary was close, as will be commented below.<br />

For a given concentration <strong>of</strong> copolymer, the value <strong>of</strong> G′ decreased<br />

with increasing temperature (Figure 11a), which is associated<br />

with the water becoming a worse solvent for poly(oxyethylene),<br />

resulting in a compression <strong>of</strong> the E-block corona <strong>and</strong>, thereby,<br />

a decrease in the effective volume fraction <strong>of</strong> micelles in the<br />

solution. The break in the curve <strong>of</strong> Figure 11a observed around<br />

35 °C is related to solution-phase separation.<br />

Frequency Dependence <strong>of</strong> Copolymer E10S10E10. The frequency<br />

dependence <strong>of</strong> the modulus was determined for solutions <strong>of</strong><br />

copolymer E10S10E10 in the concentration range 5-40% (w/v)<br />

at 25 °C. Solutions were mobile fluids over the frequency range<br />

<strong>of</strong> our measurements. Typically, values <strong>of</strong> G″ exceeded those <strong>of</strong><br />

G′ over the accessible frequency range except at low frequencies<br />

(Figure 11b). The viscoelasticity <strong>of</strong> the solution was modeled by<br />

parallel Maxwell elements:<br />

G ′ ) (G ∞ τ 2 ω 2 )/(1 + τ 2 ω 2 ) (9a)<br />

G ″ ) (G ″ τω)/(1 + τ 2 ω 2 ) (9b)<br />

where G∞ is the plateau value <strong>of</strong> G′ at high frequency, τ is<br />

the relaxation time, <strong>and</strong> ω ) 2πf (f in hertz). A good fit required<br />

the sum <strong>of</strong> several elements, but as illustrated in Figure 11b, the<br />

essential features <strong>of</strong> the rheology <strong>of</strong> the system could be<br />

reproduced using three elements, with characteristic relaxation<br />

times in the order <strong>of</strong> 0.02-0.03, 0.2-0.4, <strong>and</strong> 10-15 ms,<br />

respectively, for all the copolymer concentrations analyzed. The<br />

smallest time represents a fast process characteristic <strong>of</strong> a mobile<br />

fluid, whereas the slowest is assigned to an effect <strong>of</strong> intermicellar<br />

interaction between elongated micelles. The slight upturn in the<br />

scan at low frequency provided evidence <strong>of</strong> an emerging slow<br />

110


<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> OCH2CH(C6H5)-OCH2CH2 Block Copolymers Langmuir, Vol. 24, No. 14, 2008 7115<br />

process, which we assign to the first stages <strong>of</strong> formation <strong>of</strong> an<br />

elastic network, <strong>and</strong> which gives rise to the formation <strong>of</strong> the s<strong>of</strong>t<br />

gel. 13<br />

Discussion<br />

As commented previously, it is assumed that the parameters<br />

that control the morphologies <strong>of</strong> the aggregates are mainly core<br />

chain stretching, interfacial tension, <strong>and</strong> corona repulsion.<br />

Consequently, factors that influence the above balance can be<br />

used to control the aggregate architecture: polymer molecular<br />

weight, relative block lengths, hydrophobic block structure,<br />

solution conditions, <strong>and</strong> so forth. Compared to the separated<br />

aggregation behavior <strong>of</strong> small <strong>and</strong> large colloidal spheres, micelles<br />

<strong>and</strong> vesicles may form in dilute solution <strong>of</strong> block copolymers. 45,46<br />

In the classical <strong>and</strong> most simple description, the factor determining<br />

the shape <strong>of</strong> self-assembled amphiphilic structures is the size <strong>of</strong><br />

the hydrophobic moiety to the hydrophilic part. In particular, the<br />

micellar structure formation can be described in terms <strong>of</strong> a critical<br />

packing factor, φ, defined as<br />

φ ) V<br />

(10)<br />

al<br />

where V is the hydrophobic volume in the micelle core, a is the<br />

cross-sectional area <strong>of</strong> the hydrophilic part, <strong>and</strong> l is the chain<br />

length. 47 This model hypothesizes that changes in the amphiphile<br />

geometry, reflected in changes in V, a, orl, affect the type <strong>of</strong><br />

aggregate the amphiphile may form. At a φ value <strong>of</strong> 0.5, the<br />

model predicts formation <strong>of</strong> spherical lamellae. As the packing<br />

factor decreases, the extent <strong>of</strong> curvature increases, <strong>and</strong> the structure<br />

shape moves toward spherical micelle. Thus, in this work we<br />

prepare three different styrene oxide block copolymers with<br />

different relative block lengths <strong>and</strong> hydrophobe molecular<br />

volumes made by styrene oxide units. Since previously studied<br />

E11B11 <strong>and</strong> E14B10 were shown to form vesicles, 18 <strong>and</strong> provided<br />

that styrene oxide chain possess a larger volume than propylene<br />

oxide <strong>and</strong> butylene oxide units, we thought that copolymer E12S10<br />

should possess a hydrophilic/hydrophobic block ratio <strong>and</strong> a<br />

sufficient hydrophobe volume (because <strong>of</strong> the presence <strong>of</strong> an<br />

aromatic ring in its structure) to form vesicles, as has been<br />

corroborated. In addition, the small size <strong>of</strong> the hydrophilic block<br />

should also favor vesicle formation provided that this is favored<br />

when the hydrophilic surface area (in our case the hydrophilic<br />

corona) decreases, as has been previously observed for structurally<br />

related poly(ethylene oxide)-poly(propylene oxide) <strong>and</strong> poly-<br />

(ethylene oxide)-poly(butylene oxide) block copolymers. 17,18 On<br />

the other h<strong>and</strong>, structurally related block copolymer E11B8 (B )<br />

butylene oxide unit) 11 has been pointed out to undergo micelle<br />

elongation above 35 °C because <strong>of</strong> a severe increase in Nw <strong>and</strong><br />

geometrical considerations. The same behavior has been claimed<br />

for copolymer E17S10, for which 20 °C was enough to greatly<br />

increase Nw. 19 Nevertheless, image or other suitable techniques<br />

were not used to detect the presence <strong>of</strong> other possible structures<br />

in both cases, mainly vesicles in contrast to copolymers E11B11<br />

<strong>and</strong> E14B10 for which spontaneous vesicle formation has been<br />

demonstrated. 18<br />

On the other h<strong>and</strong>, the previous arguments would also explain<br />

why we did not find vesicular formation for the triblock E10S10E10:<br />

(46) Yu, M.; Wang, H.; Zhou, X.; Yuan, P.; Yu, C. J. Am. Chem. Soc. 2007,<br />

129, 14576–14577.<br />

(47) Israelachvili, J. N. Intermolecular <strong>and</strong> Surface Forces; Academic Press:<br />

New York; 1992; pp 380–384.<br />

(48) Kelarakis, A.; Mai, S.-M.; Havredaki, V.; Nace, V. M.; Booth, C. Phys.<br />

Chem. Chem. Phys. 2001, 3, 4037–4043.<br />

(49) Castro, E.; Taboada, P.; Mosquera, V. J. Phys. Chem. B 2006, 110, 13113–<br />

13123.<br />

this copolymer would behave as a diblock E10S5 because <strong>of</strong> the<br />

two core junctions in its structure, <strong>and</strong> not enough core volume<br />

<strong>and</strong> hydrophobic/hydrophilic ratio would be achieved to allow<br />

vesicle formation. Nevertheless, for this copolymer we found<br />

that its aggregates possess a spheroidal form, as seen by TEM,<br />

whereas for copolymer E137S18E137 micellar spheres were<br />

observed. In fact, taking the length <strong>of</strong> an S unit to be 0.36 nm<br />

per chain unit, 33 the average length <strong>of</strong> the fully stretched S10<br />

block is approximately 3.6 nm. As the central S block is looped<br />

in the micelle core for the triblock copolymer E10S10E10, the<br />

effective length will be 1.8 nm. Consequently, this would be the<br />

maximum possible radius for an anhydrous core <strong>of</strong> a spherical<br />

micelle. This means that the maximum core volume would be<br />

25 nm 3 <strong>and</strong>, given a liquid density at 15 °C<strong>of</strong>1.15gcm -3 , then<br />

the maximum aggregation number compatible with a spherical<br />

compact geometry would be ca. 15. Despite considering that (a)<br />

the S blocks possess Poisson distributions, as expected for an<br />

ideal polymerization <strong>of</strong> an alkylene oxide, 33 (b) the copolymer<br />

micelles possess water in the core associated with the hydroxyl<br />

end groups <strong>of</strong> the S blocks, as demonstrated recently, 48,49 <strong>and</strong><br />

(c) the larger entropy penalty due to chain stretching, all <strong>of</strong> which<br />

should alleviate further the stretching restriction, the derived Nw<br />

value from SLS for these copolymer (Nw ) 202) is incompatible<br />

with spherical micelles, which should be elongated, either<br />

spheroidal or wormlike. This is not the case for copolymer<br />

E137S18E137, for which a maximum aggregation number <strong>of</strong> 71 is<br />

expected to be compatible with a spherical shape. In addition,<br />

the longer E blocks increase the hydrophilic surface area<br />

promoting spherical aggregates, as commented above. The<br />

difference in micelle geometry between copolymer E137S18E137<br />

<strong>and</strong> E10S10E10 also provides an explanation for the different<br />

concentration dependences <strong>of</strong> the hydrodynamic radii <strong>and</strong> Debye<br />

functions illustrated in Figures 5 <strong>and</strong> 6, respectively. In this<br />

regard, the Debye function found for copolymer E137S18E137 is<br />

typical <strong>of</strong> spherical copolymer micelles, with the curvature being<br />

a result <strong>of</strong> interference from micelles reducing the scattering<br />

intensity as concentration is increased. However, mass action<br />

ensures that elongated micelles <strong>of</strong> copolymer E10S10E10 will<br />

increase in length as concentration is increased at a given<br />

temperature, an effect that progressively increases the scattering<br />

intensity <strong>and</strong> thus opposes the effect <strong>of</strong> micelle interaction. On<br />

the other h<strong>and</strong>, association numbers found for structurally related<br />

block copolymers with butylene oxide <strong>and</strong> styrene oxide <strong>and</strong><br />

similar hydrophobic block length shows that, for example,<br />

copolymers E13B10E13 50 (commercially available from Dow<br />

Chemical Co.) <strong>and</strong> E16B10E16 51 do not form elongated micelles<br />

in the temperature range analyzed (up to 45 °C) since their<br />

aggregation numbers are sensibly lower than that obtained for<br />

copolymer E10S10E10. Assuming a ratio <strong>of</strong> association numbers<br />

between poly(oxyethylene)-poly(oxystyrene) <strong>and</strong> poly(oxyethyelene)-poly(oxybutylene)<br />

copolymers <strong>of</strong> 3 (Nw,S/Nw,B ≈<br />

3:1), 14 the association number 51 for E16B10E16 would involve<br />

one <strong>of</strong> 153 for E10S10E10, values well outside the range possible<br />

for spherical micelles, <strong>and</strong> smaller than the value found<br />

experimentally in the present work at 15 °C. In contrast, copolymer<br />

E135B20E135 showed the formation <strong>of</strong> spherical micelles with an<br />

aggregation number <strong>of</strong> 17 at 25 °C, 22 as in the case <strong>of</strong> the closely<br />

related copolymer E137S18E137 studied in the present work.<br />

Finally, it has been shown 52 that a unifying rule for formation<br />

<strong>of</strong> polymersomes in water involves the necessity <strong>of</strong> a hydrophilicto-total<br />

mass ratio 25% < fhydrophilic < 35%. Cylindrical shapes<br />

(50) Yu, G.-E.; Yang, Y.-W.; Yang, Z.; Attwood, D.; Booth, C. Langmuir<br />

1996, 12, 3404–3412.<br />

(51) Yu, G.-E.; Li, H.; Price, C.; Booth, C. Langmuir 2002, 18, 7756–7758.<br />

(52) Discher, D. E.; Eisenberg, A. Science 2002, 297, 967–973.<br />

111


7116 Langmuir, Vol. 24, No. 14, 2008 Juárez et al.<br />

can be expected for 35% < fhydrophilic < 50%, spherical micelles<br />

for fhydrophilic < 45%, <strong>and</strong> inverted microstructures for fhydrophilic<br />

< 25%. Although sensitivities <strong>of</strong> these rules to chain chemistry<br />

<strong>and</strong> molecular weight have not been fully probed, the present<br />

analyzed copolymers seem to be in agreement with these.<br />

Conclusions<br />

We have analyzed the micellization, gelation, <strong>and</strong> structure<br />

<strong>of</strong> the aggregates <strong>of</strong> three poly(ethylene oxide)-polystyrene oxide<br />

block copolymers. Two <strong>of</strong> them have similar block length but<br />

different structures (E12S10,E10S10E10), <strong>and</strong> the other has longer<br />

blocks (E137S18E137). In this way, we can compare the effect <strong>of</strong><br />

temperature <strong>and</strong> block length on the self-assembly properties.<br />

The values <strong>of</strong> the critical micelle concentration show that the<br />

short triblock copolymer has a cmc twice as large as that <strong>of</strong> the<br />

related diblock. The effect is originated by entropy <strong>of</strong> the triblock<br />

chains constrained by two block junctions in the core interface<br />

<strong>of</strong> the aggregate, compared with only one constraint for the diblock<br />

chain. Intensity fraction size distributions showed the existence<br />

<strong>of</strong> several species in solutions <strong>of</strong> copolymer E12S10 whereas only<br />

one for copolymers E10S10E10 <strong>and</strong> E137S18E137. Light scattering,<br />

polarized light microscopy, <strong>and</strong> transmission <strong>and</strong> cryo-scanning<br />

electron microscopies provided sufficient evidence to ensure the<br />

coexistence <strong>of</strong> micelles <strong>and</strong> polydisperse vesicles for the most<br />

concentrated samples <strong>of</strong> this copolymer (but even in the dilute<br />

regime). The size <strong>of</strong> these vesicular structures varies between 60<br />

<strong>and</strong> 500 nm as shown by TEM, DLS, <strong>and</strong> cryo-SEM. The classical<br />

birefringent pattern <strong>of</strong> vesicles was observed by polarized light<br />

microscopy. On the other h<strong>and</strong>, the single size distribution<br />

observed for copolymers E10S10E10 <strong>and</strong> E137S18E137 was assigned<br />

to elongated <strong>and</strong> spherical micelles, respectively. This assignment<br />

was made in light <strong>of</strong> the high aggregation number derived for<br />

the former copolymer which is incompatible with a spherical<br />

geometry <strong>of</strong> the aggregated, as also corroborated by TEM. Tube<br />

inversion was used to define the mobile-immobile (s<strong>of</strong>t-hard<br />

gel) phase boundaries. The phase diagram <strong>of</strong> copolymer<br />

E137S18E137 was further completed to determine the sol-s<strong>of</strong>t<br />

gel-hard gel boundaries. In contrast, copolymers E12S10 <strong>and</strong><br />

E10S10E10 did not gel in the concentration range analyzed. Only<br />

certain concentrations <strong>of</strong> copolymer E10S10E10 were analyzed by<br />

rheometry, for which an upturn in the low-frequency range <strong>of</strong><br />

the stress moduli was observed, denoting an evidence <strong>of</strong> an<br />

emerging slow process, which we assign to the first stages <strong>of</strong><br />

formation <strong>of</strong> an elastic network.<br />

Acknowledgment. We thank the Ministerio de Educación y<br />

Ciencia through projects MAT2007-61604 <strong>and</strong> Xunta de Galicia<br />

for financial support. J.J. thanks USC for his PhD fellowship,<br />

<strong>and</strong> E.C. thanks Xunta de Galicia for his “Anxeles Alvariño”<br />

research contract. We are really grateful to Pr<strong>of</strong>s. D. Attwood<br />

<strong>and</strong> C. Booth for providing several <strong>of</strong> the copolymer samples.<br />

LA8004568<br />

112


Surface Properties <strong>of</strong> Monolayers <strong>of</strong> Amphiphilic Poly(ethylene oxide)-Poly(styrene oxide)<br />

Block Copolymers<br />

Introduction<br />

Josué Juárez, † Sonia Goy-López, † Adriana Cambón, † Miguel A. Valdez, ‡ Pablo Taboada,* ,† <strong>and</strong><br />

Víctor Mosquera †<br />

Grupo de Física de Coloides y Polímeros, Departamento de Física de la Materia Condensada, Facultad de<br />

Física, UniVersidad de Santiago de Compostela, E-15782, Santiago de Compostela, Spain, <strong>and</strong> Departamento<br />

de InVestigación en Polímeros y Materiales y Departamento de Física, UniVersidad de Sonora, Resales y<br />

TransVersal, 83000 Hermosillo Sonora, Mexico<br />

ReceiVed: May 31, 2010; ReVised Manuscript ReceiVed: July 30, 2010<br />

The surface properties <strong>of</strong> poly(styrene oxide)-poly(ethylene oxide) block copolymers EO12SO10,EO10SO10EO10,<br />

<strong>and</strong> EO137SO18EO137 at the air-water (a/w) <strong>and</strong> chlor<strong>of</strong>orm-water (c/w) interfaces have been analyzed by<br />

Tracker drop tensiometry, Langmuir film balance, <strong>and</strong> atomic force microscopy (AFM). The kinetic adsorptions<br />

for the block copolymer solutions at both interfaces were determined by the pendant drop technique. At the<br />

a/w, the polymer adsorption is reduced as the polymer hydrophobicity increases. Measurements <strong>of</strong> the interfacial<br />

rheological behavior <strong>of</strong> the copolymers showed that the adsorption layers at the a/w interface manifest obvious<br />

solid-like properties in the whole accessible frequency range, whereas a viscous fluid-like behavior is displayed<br />

at the c/w interface. The differences observed in the monolayer isotherms obtained for the three copolymers<br />

arose from their different hydrophobic/hydrophilic block ratios <strong>and</strong> block lengths. In this regard, copolymer<br />

EO137SO18EO137 displays an adsorption isotherm with the four classical regions (pancake, mushroom, brush,<br />

<strong>and</strong> condensed states), whereas for EO12SO10 <strong>and</strong> EO10SO10EO10 copolymers, only two regions were observed.<br />

The topographic images <strong>of</strong> copolymer films were obtained by AFM in noncontact mode. Surface circular<br />

micelles are observed at the two surface transfer pressures studied. A micelle size decrease <strong>and</strong> an increase<br />

in monolayer thickness are observed with an increase in transfer pressure. At the largest transfer pressure<br />

used, elongated aggregates were observed. Aggregation numbers derived from AFM pictures were in fair<br />

agreement with those obtained for micelles in solution, <strong>and</strong> they became larger as the SO weight fraction<br />

increases at a certain deposition pressure. This can be a consequence <strong>of</strong> stronger attractive interactions between<br />

the SO blocks to avoid contact with the solvent.<br />

Amphiphilic block copolymers are an important class <strong>of</strong><br />

materials that have attracted considerable attention because <strong>of</strong><br />

their outst<strong>and</strong>ing solution properties, such as their self-assembly<br />

in the presence <strong>of</strong> a selective solvent or surface. 1 In particular,<br />

the interfacial properties <strong>of</strong> block copolymers have received<br />

great interest in the last years due to their role in the industrial<br />

<strong>and</strong> biological fields. 2-4 The ability <strong>of</strong> block copolymers to<br />

adsorb at an interface, modulate their properties, <strong>and</strong> even to<br />

self-assemble into well-defined nanoscale structures in two<br />

dimensions has led to their use in colloidal stability <strong>and</strong> coating<br />

processes, 5 as steric barriers at solid surfaces in medical devices<br />

to avoid undesired protein adhesion, 6 as stabilizers <strong>of</strong> liquid<br />

interfaces <strong>of</strong> foams 7 <strong>and</strong> emulsions 8 to prevent coalescence <strong>and</strong><br />

flocculation, as templates for the rational design <strong>of</strong> large-scale<br />

nanometer architectures in thin film configurations, 9,10 <strong>and</strong> as<br />

precursors for nanoporous materials. 11<br />

The self-assembly behavior <strong>and</strong> physicochemical properties<br />

<strong>of</strong> diblock <strong>and</strong> triblock copolymers <strong>of</strong> polyoxyalkylenes in<br />

solution have been extensively studied. 12-21 Variation <strong>of</strong><br />

hydrophobic monomer, block length, <strong>and</strong> architecture allows<br />

close control <strong>of</strong> their physicochemical properties. The most<br />

* To whom correspondence should be addressed. E-mail: pablo.taboada@<br />

usc.es. Tel: 0034981563100, ext 14042. Fax: 0034981520676.<br />

† Universidad de Santiago de Compostela.<br />

‡ Universidad de Sonora.<br />

J. Phys. Chem. C 2010, 114, 15703–15712 15703<br />

10.1021/jp1049777 © 2010 American Chemical Society<br />

Published on Web 08/26/2010<br />

familiar copolymers <strong>of</strong> this type are those whose hydrophobic<br />

block is formed by units <strong>of</strong> oxypropylene (-OCH2CH(CH3),<br />

denoted as PO), 1,2-oxybutylene (-OCH2CH(C2H5), denoted<br />

as BO), <strong>and</strong> oxyphenylethylene (-OCH2CH-(C6H5)), prepared<br />

from styrene oxide (denoted as SO). These copolymers are<br />

already commercially available from BASF, 22 Dow Chemical<br />

Co., 23 <strong>and</strong> Goldschimdt AG, 24 respectively. Nevertheless, the<br />

two-dimensional properties <strong>and</strong> aggregation <strong>of</strong> these classes <strong>of</strong><br />

copolymers at the air/aqueous interface has been only extensively<br />

reported for poly(ethylene oxide)-poly(propylene oxide)poly(ethylene<br />

oxide) (EO-PO-EO) block copolymers or Pluronics,<br />

the X-star shaped EO-PO-EO copolymers known as<br />

Tetronic copolymers (from BASF), 27,28 <strong>and</strong> poly(butylene<br />

oxide)-poly(ethylene oxide) block copolymers (BO-EO). 29 Very<br />

little or no report has been put into fundamentally underst<strong>and</strong>ing<br />

the 2-D dimensional behavior <strong>of</strong> poly(ethyelene oxide)poly(styrene<br />

oxide) block copolymers at interfaces. Therefore,<br />

fundamental studies <strong>of</strong> SO-based block copolymers performed<br />

at either the air/water or the air/organic solvent interface can<br />

provide valuable information on their interfacial phase behavior<br />

to further guide their possible use in different biomedical <strong>and</strong><br />

coating applications. To fill this gap, in the present work, we<br />

studied the surface behavior <strong>and</strong> surface properties <strong>of</strong> three<br />

poly(ethylene oxide)-poly(styrene oxide) block copolymers,<br />

EO12SO10, EO10SO10EO10, <strong>and</strong> EO137SO18EO137 (where the<br />

subscripts denote the respective block length) by dynamic<br />

113


15704 J. Phys. Chem. C, Vol. 114, No. 37, 2010 Juárez et al.<br />

TABLE 1: Molecular Characteristics <strong>of</strong> the Copolymers<br />

Mn a /g mol -1<br />

(NMR)<br />

wt%SaMw/Mn (NMR)<br />

a Mw/g<br />

(GPC) mol-1 cmcb /g<br />

dm-3 Γb max/10-6 mol m-2 EO12SO10 1660 73 1.05 1760 0.006 0.97<br />

EO10SO10EO10 1980 55 1.06 2130 0.012 0.91<br />

EO137SO18EO137 14 200 15 1.06 15 100 0.017 0.40<br />

a Estimated uncertainty: Mn to (3%;wt%Sto(1%, Mw/Mn to<br />

(0.01. Mw calculated from Mn <strong>and</strong> Mw/Mn. b Data from ref 32 at 15<br />

°C.<br />

interfacial tension, dilatational rheology, Langmuir monolayers,<br />

<strong>and</strong> atomic force microscopy (AFM). The self-assembling<br />

properties <strong>and</strong> solubilization capacity <strong>of</strong> these copolymers in<br />

bulk aqueous solution were analyzed in recent studies, 30-32 in<br />

which the existence <strong>of</strong> vesicular structures (polymersomes),<br />

elongated micelles, <strong>and</strong> spherical micelles in bulk solution was<br />

reported for copolymers EO12SO10, EO10SO10EO10, <strong>and</strong><br />

EO137SO18EO137, respectively. These block copolymers display<br />

different volume-to-length ratios <strong>of</strong> the hydrophobic block,<br />

relative lengths <strong>of</strong> the polyoxyethylene blocks, <strong>and</strong> polymer<br />

architectures, which enable analyzing the influence <strong>of</strong> these<br />

parameters on the block copolymer surface behavior. In this<br />

regard, due to its lengthy EO blocks <strong>and</strong> higher hydrophilicity<br />

(as denoted by its larger cmc values; see Table 1), copolymer<br />

EO137SO18EO137 displays a clear <strong>and</strong> complete adsorption<br />

isotherm in which the different conformational states <strong>of</strong> the<br />

polymer chains (“pancake”, “mushroom”, <strong>and</strong> “brush” conformations)<br />

are observed, with a significant viscous fluid behavior<br />

<strong>and</strong> a spherical morphology <strong>of</strong> its aggregates at the surface film.<br />

In contrast, the surface pressures <strong>of</strong> EO12SO10 <strong>and</strong> EO10SO10EO10<br />

monolayers were measurable at a significantly smaller area per<br />

molecule, <strong>and</strong> only two different clear regions can be distinguished<br />

in their isotherms as a consequence <strong>of</strong> the reduction <strong>of</strong><br />

the fractional interfacial area occupied by EO segments. In<br />

addition, copolymer EO12SO10 displays a clear rheological solidlike<br />

behavior <strong>and</strong> its organization in elongated structures similar<br />

to a string <strong>of</strong> beads upon self-assembly at the air-water (a/w)<br />

interface. Finally, a comparison <strong>of</strong> the surface behavior <strong>of</strong> these<br />

copolymers with that observed for structurally related polystyrenepoly(ethylene<br />

oxide) (PS-PEO) block copolymers is also<br />

established. 33,34<br />

Experimental Section<br />

Materials. Copolymers EO12SO10, EO10SO10EO10, <strong>and</strong><br />

EO137SO18EO137 were prepared by sequential oxyanionic polymerization<br />

<strong>of</strong> styrene oxide, followed by ethylene oxide,<br />

starting from a mon<strong>of</strong>unctional or difunctional initiator to form<br />

the hydrophobic block as described in detail by Crothers et al. 35<br />

Table 1 shows the molecular characteristics, critical micellar<br />

concentrations (cmc), <strong>and</strong> surface excess concentrations (Γmax)<br />

<strong>of</strong> the copolymers. Water was double-distilled <strong>and</strong> degassed<br />

before used.<br />

Dynamic Interfacial Tension <strong>and</strong> Dilatational Rheology<br />

Measurements. Pendant bubble tensiometry was used to<br />

determine the dynamic interfacial tension <strong>and</strong> the interfacial<br />

dilatational rheology at two different interfaces: air-water<br />

(a/w) <strong>and</strong> chlor<strong>of</strong>orm-water (c/w) interfaces. The bubble was<br />

formed at the tip <strong>of</strong> a U-shaped stainless steel needle (0.5 mm<br />

inner diameter) immersed in an aqueous block copolymer<br />

solution <strong>of</strong> the desired concentration, with the air bubble upward<br />

for the (a/w) interface, whereas downward for the (c/w)<br />

interface. Measurements were carried out in a Track tensiometer<br />

equipment (I.T. Concept, France) adapted to determine surface<br />

tension values in real time with an accuracy <strong>of</strong> (0.1 mN/m.<br />

Interfacial tension <strong>and</strong> interfacial rheology estimations are based<br />

on the digital pr<strong>of</strong>ile <strong>of</strong> a drop image <strong>and</strong> the resolution <strong>of</strong> the<br />

Gauss-Laplace equation. Win Drop s<strong>of</strong>tware (I.T. Concept,<br />

France) was used to obtain the surface tension values by means<br />

<strong>of</strong> the axisymetric drop shape analysis.<br />

The aqueous solutions concentrations <strong>of</strong> the EO12SO10,<br />

EO10SO10EO10, <strong>and</strong> EO137SO18EO137 block copolymers were<br />

prepared at a constant concentration <strong>of</strong> 2 × 10 -3 mg/mL, below<br />

their respective cmc values. At the (a/w) interface, the analysis<br />

<strong>of</strong> the interfacial tension (γ) measurements was followed in a<br />

relative short time (2000 s) after the bubble was formed, <strong>and</strong><br />

the dilatational elastic storage (E′) <strong>and</strong> loss (E′′) moduli were<br />

determined after 2000 s using a frequency sweep <strong>of</strong> 3.14, 2.09,<br />

1.04, 0.52, 0.31, 0.15, 0.078, <strong>and</strong> 0.039 rad/s. All measurements<br />

were carried out in a time interval <strong>of</strong> 400 s using a 10%<br />

amplitude oscillation <strong>of</strong> the maximum volume drop. On the other<br />

h<strong>and</strong>, experimental data <strong>of</strong> interfacial tension, γ, at the<br />

chlor<strong>of</strong>orm-water interface were followed for 180 s, <strong>and</strong> after<br />

this time, volume oscillations to the stabilized pendant drop<br />

volume were applied. Storage, E′, <strong>and</strong> loss, E”, moduli were<br />

determined in this way.<br />

Briefly, the analysis <strong>of</strong> the viscoelastic properties <strong>of</strong> the<br />

absorbed layer was performed by means <strong>of</strong> a sinusoidal<br />

deformation <strong>of</strong> the drop area, A, at a defined frequency, ω,<br />

produced by the Tracker tensiometer at a time t. Calculus <strong>of</strong><br />

the complex dilatational elasticity modulus E(ω) is determined<br />

by<br />

E(ω) )-A dy<br />

dA<br />

This can also be defined as a complex function, which can be<br />

written as<br />

where E′ <strong>and</strong> E′′ correspond to the storage <strong>and</strong> the loss moduli,<br />

respectively, by analogy with 3-D rheology. Assuming that the<br />

mechanical properties <strong>of</strong> the adsorbed layer follow a Maxwell<br />

model behavior, the storage <strong>and</strong> loss moduli can be written as<br />

follows<br />

where E0i <strong>and</strong> ω0i (ω0i ) 2π/τ0i) are the corresponding weights<br />

<strong>of</strong> the ith relaxation element to the total elasticity modulus E0<br />

<strong>and</strong> the characteristic relaxation frequency, respectively; τ0i is<br />

the characteristic relaxation time <strong>of</strong> the Maxwell model.<br />

Adsorption Isotherms. Surface-pressure isotherms were<br />

recorded for monolayers spread from chlor<strong>of</strong>orm solutions<br />

(1 mg/mL) onto a nanopure-quality water subphase in a Langmuir-Blodgett<br />

(LB) Teflon trough (model 611 from Nima<br />

Technologies Ltc., Coventry, U.K.). Volumes ranging from 10<br />

to 50 µL <strong>of</strong> either EO12SO10,EO10SO10EO10,orEO137SO18EO137<br />

block copolymer solutions (1 mg/mL) were spread dropwise<br />

on a Millipore water subphase with a Hamilton microsyringe.<br />

To ensure complete evaporation <strong>of</strong> the solvent, a time lag <strong>of</strong><br />

(1)<br />

E(ω) ) E'(ω) + iE''(ω) (2)<br />

E'(ω) ) ∑ i<br />

E''(ω) ) ∑ i<br />

(ω/ω0i )<br />

E0i 2<br />

1 + (ω/ω0i ) 2<br />

(ω/ω0i )<br />

E0i 1 + (ω/ω0i ) 2<br />

(3)<br />

(4)<br />

114


Amphiphilic PEO-PSO Block Copolymers J. Phys. Chem. C, Vol. 114, No. 37, 2010 15705<br />

15 min was applied between the deposition <strong>of</strong> the copolymer<br />

<strong>and</strong> the beginning <strong>of</strong> compression <strong>of</strong> the spread molecules. After<br />

chlor<strong>of</strong>orm was evaporated, the surface area was compressed<br />

by Teflon barriers at a rate <strong>of</strong> 5 mm · min -1 . The temperature<br />

was kept constant at 25 °C with a water bath, <strong>and</strong> all<br />

experiments were carried out inside a dust-free glass box. The<br />

surface pressure was measured using a Platinum Wilhelmy plate<br />

located midway between the two barriers <strong>and</strong> oriented perpendicular<br />

to them, that is, parallel to the direction <strong>of</strong> the barrier<br />

movement. The complete system was operated on an antivibration<br />

table. On the other h<strong>and</strong>, Langmuir films were transferred<br />

onto a freshly cleaved mica substrate for imaging. After<br />

spreading the chlor<strong>of</strong>orm solutions, again at least 15 min was<br />

allowed for solvent evaporation. The barriers were then<br />

compressed to the targeted surface pressures, 5 <strong>and</strong> 11 mN/m,<br />

respectively, at a speed <strong>of</strong> 5 mm/min. Once the targeted surface<br />

pressure was reached <strong>and</strong> stable, the cleaved mica was lifted<br />

up at a speed <strong>of</strong> 1 mm/min. Film transfer ratios <strong>of</strong> 0.95-1.0<br />

were obtained. The resulting films were individually stored at<br />

room temperature in clean closed vials.<br />

Atomic Force Microscopy. AFM images <strong>of</strong> block copolymer<br />

monolayers were performed on freshly cleaved mica substrates.<br />

The measurements were performed in a JEOL instrument (model<br />

JSPM 4210) in noncontact mode using nitride cantilevers<br />

NSC15 from MicroMasch, U.S.A. (typical working frequency<br />

<strong>and</strong> spring constant <strong>of</strong> 325 kHz <strong>and</strong> 40 N/m, respectively). The<br />

AFM samples were dried in air or under a nitrogen flow when<br />

required. Control samples (freshly cleaved mica <strong>and</strong> buffer<br />

solution) were also investigated to exclude possible artifacts.<br />

Topography <strong>and</strong> phase-shift data were collected in the trace <strong>and</strong><br />

retrace direction <strong>of</strong> the raster, respectively. The <strong>of</strong>fset point was<br />

adapted accordingly to the roughness <strong>of</strong> the sample. The scan<br />

size was usually 500 nm (aspect ratio, 1 × 1), with a sample<br />

line <strong>of</strong> 256 points <strong>and</strong> a step size <strong>of</strong> 1 µm. The scan rate was<br />

tuned proportionally to the area scanned <strong>and</strong> kept within the<br />

0.35-2 Hz range. Each sample was imaged several times at<br />

different locations on the substrate to ensure reproducibility. In<br />

all cases, the imaging <strong>of</strong> LB films was performed far from the<br />

edge <strong>of</strong> the glass substrate to minimize local effects caused by<br />

turbulent water flow at the boundary <strong>and</strong> meniscus effects during<br />

transfer. Diameters <strong>and</strong> heights <strong>of</strong> copolymer aggregates were<br />

determined by sectional analysis taken from the average <strong>of</strong><br />

several sections through the aggregates.<br />

Results <strong>and</strong> Discussion<br />

Some molecular characteristics <strong>of</strong> the investigated block<br />

copolymers EO12SO10,EO10SO10EO10, <strong>and</strong> EO137SO18EO137 are<br />

shown in Table 1. As expected, the two triblock copolymers<br />

have higher cmc values than the diblock. This effect originates<br />

from the entropy <strong>of</strong> the triblock chains constrained by two block<br />

junctions in the core-shell interface <strong>of</strong> the micelle compared<br />

to the only constrain for the diblock chain. The copolymers were<br />

chosen to form two pairs, one with similar-sized hydrophobic<br />

SO blocks but different polymer architectures (EO12SO10 <strong>and</strong><br />

EO10SO10EO10) <strong>and</strong> the other with completely different hydrophilic<br />

EO block lengths (EO10SO10EO10 <strong>and</strong> EO137SO18EO137)<br />

<strong>and</strong> relatively similar hydrophobic SO blocks. This was made<br />

with the purpose <strong>of</strong> comparing the relative importance <strong>of</strong> the<br />

polymer architecture <strong>and</strong> the hydrophilic/hydrophobic block ratio<br />

on both polymer adsorption <strong>and</strong> surface rheological behavior<br />

at the air/water <strong>and</strong> chlor<strong>of</strong>orm/water interfaces. In previous<br />

studies, the polymer architecture, the volume-to-length ratios<br />

<strong>of</strong> the hydrophobic block, <strong>and</strong> the relative lengths <strong>of</strong> the<br />

polyoxyethylene blocks were found to be helpful in controlling<br />

the micellar morphology <strong>of</strong> these copolymers. 32,36 In particular,<br />

the coexistence <strong>of</strong> spherical micelles with vesicles (polymersomes)<br />

was observed for copolymer EO12SO10; in contrast, the<br />

formation <strong>of</strong> elongated <strong>and</strong> spherical micelles for copolymers<br />

EO10SO10EO10 <strong>and</strong> EO137SO18EO137, respectively, was found<br />

upon their self-assembly. 32<br />

Dynamic Surface Tension<br />

The surface tension <strong>of</strong> the copolymers as a function <strong>of</strong> time<br />

for copolymers EO12SO10,EO10SO10EO10, <strong>and</strong> EO137SO18EO137<br />

at a concentration <strong>of</strong> 2 × 10-3 mg/mL is illustrated in Figure<br />

1. The low concentration used allowed us to work in relative<br />

short time intervals to achieve a quasi-stabilization at both<br />

air-water (a/w) <strong>and</strong> chlor<strong>of</strong>orm-water (c/w) interfaces, <strong>and</strong> it<br />

also avoided complications associated with chlor<strong>of</strong>orm evaporation.<br />

Nevertheless, the true equilibrium state was not attained<br />

for any <strong>of</strong> the copolymers within 1h<strong>of</strong>adsorption because the<br />

surface tension still continued to decrease slightly with time.<br />

Usually, the dynamic surface tension curves can be divided<br />

in three stages: At relative low polymer concentrations, a quasiplateau<br />

or a very slow decrease <strong>of</strong> γ(t) can be distinguished at<br />

short times, which corresponds to the so-called lag phase. The<br />

postlag phase stage <strong>of</strong> the polymer adsorption process is<br />

characterized by a rapid decrease <strong>of</strong> γ(t), which is associated<br />

with the increased number <strong>of</strong> copolymer molecules adsorbed<br />

at the interface <strong>and</strong> to conformational changes <strong>of</strong> the polymer<br />

chains adsorbed at the interface. In this stage, lateral interactions<br />

between adsorbed macromolecules stabilize the monolayer <strong>and</strong><br />

the adsorption <strong>of</strong> the molecules continues to be controlled by<br />

macromolecule diffusion without appreciable desorption. Finally,<br />

the postlag stage transforms to the final (or last) adsorption<br />

phase, which is recognized by some slowing down <strong>of</strong> the<br />

interfacial tension decrease rate. At this stage, the dense<br />

adsorption layer exerts electrostatic <strong>and</strong> steric repulsion toward<br />

newly arrived macromolecules <strong>and</strong> hinders their adsorption at<br />

the interface. 37,38 Nevertheless, despite their hindrance, the<br />

interfacial tension tends to continuously decrease during this<br />

stage because <strong>of</strong> the high surface activity <strong>of</strong> the block<br />

copolymers.<br />

In our case, initial γ values <strong>of</strong> 66.4, 64.5, <strong>and</strong> 57.1 mN/m<br />

were observed for copolymers EO12SO10, EO10SO10EO10, <strong>and</strong><br />

EO137SO18EO137, respectively (see Figure 1a), as a consequence<br />

<strong>of</strong> the very rapid migration <strong>of</strong> copolymer molecules from the<br />

bulk solution to the a/w interface, giving rise to a block<br />

copolymer monolayer. For copolymer EO137SO18EO137, we<br />

assume that the observed experimental data corresponds to the<br />

last part <strong>of</strong> the postlag time stage <strong>and</strong> the final polymeric<br />

adsorption stage. In contrast, for both EO12SO10 <strong>and</strong><br />

EO10SO10EO10, the curve pr<strong>of</strong>ile can be divided in three stages<br />

after an almost instantaneous surface tension decrease. We will<br />

keep the same nomenclature (lag phase, postlag phase, <strong>and</strong> final<br />

adsorption phase) to define the regions in these curves despite<br />

that their true meaning corresponds to values departing from<br />

the surface tension value <strong>of</strong> pure water. For copolymer<br />

EO12SO10, the lag phase is longer, <strong>and</strong> the decrease <strong>of</strong> γ in the<br />

postlag phase is faster than for copolymer EO10SO10EO10,<br />

probably as a result <strong>of</strong> its greater hydrophobicity. Similarly, a<br />

much larger EO block for copolymer EO137SO18EO137 if<br />

compared to EO10SO10EO10 results in a larger surface tension<br />

decrease in the very early stages <strong>of</strong> the adsorption process, which<br />

we attribute to an increased repulsion between the EO chains,<br />

giving rise to an enhancement <strong>of</strong> the surface pressure.<br />

If the adsorption kinetics is controlled by the diffusion <strong>of</strong><br />

macromolecules from the bulk solution to the interface, for<br />

115


15706 J. Phys. Chem. C, Vol. 114, No. 37, 2010 Juárez et al.<br />

Figure 1. (a) Variation <strong>of</strong> surface tension, γ, with time at the air-water<br />

interface for copolymers (O) EO12SO10, (b) EO10SO10EO10, <strong>and</strong> (2)<br />

EO137SO18EO137. (b) Variation <strong>of</strong> surface concentration, Γ, with time<br />

at the air-water interface for copolymers (s) EO12SO10, (·····)<br />

EO10SO10EO10, <strong>and</strong> (----) EO137SO18EO137. (c) Variation <strong>of</strong> surface<br />

tension, γ, with time at the chlor<strong>of</strong>orm-water interface for copolymers<br />

(9) EO12SO10, (b) EO10SO10EO10, <strong>and</strong> (2) EO137SO18EO137.<br />

strongly adsorbing species, this is described by the Ward-Torday<br />

equation, 39 Γ(t) ) 2c0(Dt/π) 1/2 , where c0 is the polymer<br />

concentration <strong>and</strong> D the bulk diffusion coefficient. A diffusioncontrolled<br />

adsorption mechanism will be expected to give a<br />

straight line in a plot <strong>of</strong> Γ versus t 1/2 . In these experiments, the<br />

direct connection between Γ <strong>and</strong> t cannot be measured, but a<br />

measured value <strong>of</strong> Π may be converted to a relative value <strong>of</strong> Γ<br />

by using the scaling expression Π ∼ Γ y <strong>and</strong> the calculated value<br />

<strong>of</strong> y from the adsorption isotherm experiments (see below).<br />

Figure 1b shows such a plot for the three copolymers. The<br />

proportionally constant was set equal to 1 (m 2 · mg -2 ) y mN · m -1 ,<br />

which implies that the surface pressure is set to 1 mN · m -1 at<br />

a surface concentration <strong>of</strong> 1 mg · m -2 . This value is typical for<br />

that found for many polymers. 40,41 From this figure, it can be<br />

observed that the expected straight line is quite steep at very<br />

short times, but a curvature in the plot is present at larger times.<br />

This may indicate that the diffusion-controlled adsorption<br />

process only operates during the first few seconds <strong>of</strong> the process<br />

<strong>and</strong>, then, is more likely controlled by a slow configurational<br />

change <strong>and</strong> reorientation <strong>of</strong> the polymer molecules at the surface.<br />

At this respect, a bimodal type <strong>of</strong> adsorption kinetics <strong>of</strong> Pluronic<br />

block copolymers, including initial diffusion, followed by an<br />

internal reorganization <strong>of</strong> the structure, was discussed recently. 42<br />

On the other h<strong>and</strong>, there was no measurable effect on the<br />

interfacial tension at the chlor<strong>of</strong>orm/water interface (c/w) due<br />

to the adsorption <strong>of</strong> copolymer EO12SO10 (see Figure 1c). This<br />

behavior might originate from the small amount <strong>of</strong> copolymer<br />

at the interface due to the slow diffusion <strong>of</strong> their chains to the<br />

interface in combination with a low bulk concentration. The<br />

slow diffusion can be associated with the fact that chlor<strong>of</strong>orm<br />

is a good solvent for both EO <strong>and</strong> SO blocks, <strong>and</strong> provided<br />

that copolymer EO12SO10 is very hydrophobic, the copolymer<br />

chains prefer being in the organic phase. This can be additionally<br />

corroborated with the progressively faster decrease observed in<br />

surface tension values at short times for the increasingly<br />

hydrophilic copolymers EO10SO10EO10 <strong>and</strong> EO137SO18EO137.<br />

This faster decrease also enables the development <strong>of</strong> a postlag<br />

phase for these copolymers. In fact, for the latter copolymer,<br />

the curve pr<strong>of</strong>ile is very similar to that obtained at the a/w<br />

interface, although the absolute γ values are quite different as<br />

a result <strong>of</strong> the depression <strong>of</strong> interfacial tension due to the<br />

presence <strong>of</strong> chlor<strong>of</strong>orm.<br />

Dilatational Rheology. Analysis <strong>of</strong> the frequency dependence<br />

<strong>of</strong> the dilational moduli allows us to get detailed information<br />

about the type <strong>of</strong> interactions existing between polymer chains<br />

located at the interface. Figure 2 shows the frequency dependence<br />

<strong>of</strong> storage, E′, <strong>and</strong> loss, E′′, dilatational moduli in<br />

semilogarithmic plots for the adsorbed layers <strong>of</strong> block copolymers<br />

EO12SO10,EO10SO10EO10, <strong>and</strong> EO137SO18EO137 at both the<br />

a/w <strong>and</strong> the c/w interfaces. The frequency dependence <strong>of</strong> both<br />

E′ <strong>and</strong> E′′ is generally believed to be caused by relaxation<br />

processes at the interface. 43,44 A satisfactory rheological description<br />

<strong>of</strong> the adsorption layers <strong>of</strong> macromolecules may be done<br />

by considering several characteristic relaxation frequencies (or<br />

relaxationtimes,τ),whicharerelatedtotheadsorption-desorption<br />

exchange <strong>of</strong> polymer chains between the surface <strong>and</strong> the<br />

adjacent subinterface layer during dilatational perturbation <strong>and</strong><br />

molecular reconformation <strong>of</strong> the adsorbed layer. 38 In this way,<br />

the frequency dependences <strong>of</strong> storage (E′) <strong>and</strong> loss (E′′) moduli<br />

were fitted to the sum <strong>of</strong> two Maxwell elements<br />

(ω/ω01 )<br />

E'(ω) ) E01 2<br />

1 + (ω/ω01 ) 2 + E (ω/ω02 )<br />

02<br />

2<br />

1 + (ω/ω02 ) 2<br />

(ω/ω01 )<br />

E''(ω) ) E01 1 + (ω/ω01 ) 2 + E (ω/ω02 )<br />

02<br />

1 + (ω/ω02 ) 2<br />

where E′ is related to lateral interactions between polymer<br />

segments at the interface plane <strong>and</strong> is relevant for the rigidity<br />

<strong>of</strong> interfacial film, whereas E′′ is related to molecular reorga-<br />

(5)<br />

(6)<br />

116


Amphiphilic PEO-PSO Block Copolymers J. Phys. Chem. C, Vol. 114, No. 37, 2010 15707<br />

Figure 2. Storage, E′ (closed symbols), <strong>and</strong> loss, E′′ (open symbols),<br />

moduli as a function <strong>of</strong> frequency for copolymers (9) EO12SO10, (b)<br />

EO10SO10EO10, <strong>and</strong> (2) EO137SO18EO137 at the (a) air-water <strong>and</strong> (b)<br />

chlor<strong>of</strong>orm-water interfaces.<br />

nization processes, such as the expulsion <strong>of</strong> polymer chains from<br />

the interface upon compression <strong>and</strong> the interactions <strong>of</strong> polymer<br />

molecules with adjacent liquid molecules. 45 The first characteristic<br />

relaxation time, τ01, is usually attributed to the<br />

adsorption-desorption exhange <strong>of</strong> molecules <strong>and</strong> polymer<br />

segments between the surface <strong>and</strong> the adjacent subsurface layer<br />

during dilatational perturbations <strong>of</strong> the interface; τ02 corresponds<br />

to slower reconformations <strong>of</strong> the adsorbed macromolecules<br />

inside the adsorption layer.<br />

At the a/w interface, the adsorption layers manifest obvious<br />

solid-like properties in the whole accessible frequency range<br />

(ω ) 0.039-3.14 rad/s) with E′ >E′′, following the sequence<br />

EO12SO10 > EO10SO10EO10 > EO137SO18EO137, <strong>and</strong> with a low<br />

significance <strong>of</strong> dissipative processes during dilational deformations<br />

<strong>of</strong> these block copolymer layers. These values suggest a<br />

more fluid layer in the case <strong>of</strong> copolymer EO137SO18EO137 as a<br />

consequence <strong>of</strong> its lower hydrophobicity (larger cmc) due to<br />

the presence <strong>of</strong> extended oxyethyelene chains. In contrast,<br />

copolymer EO12SO10 shows the most important solid-like<br />

behavior, which involves the formation <strong>of</strong> tightly packed<br />

copolymer chains at the interface to avoid contact between<br />

hydrophobic blocks <strong>and</strong> water. Molar copolymer concentration<br />

differences slightly affect the final dilatational moduli values<br />

(not shown). We also point out that the storage modulus E′<br />

slightly, but continuously, increases with frequency, whereas<br />

TABLE 2: Characteristic Relaxation Times, τ01 <strong>and</strong> τ02, <strong>and</strong><br />

Characteristic Dilatational Moduli Weights <strong>of</strong> Each<br />

Relaxation Mode, E01 <strong>and</strong> E02, <strong>of</strong> Copolymers EO12SO10,<br />

EO10SO10EO10, <strong>and</strong> EO137SO18EO137 at the Air-Water <strong>and</strong><br />

Chlor<strong>of</strong>orm-Water Interfaces<br />

τ01 a /s τ02 a /s E01 a /mN m -1 E02 a /mN m -1<br />

EO12SO10 3.2 48.0<br />

air-water<br />

4.9 19.3<br />

EO10SO10EO10 10.0 75.0 6.8 13.8<br />

EO137SO18EO137 6.6 81.1 1.7<br />

chlor<strong>of</strong>orm-water<br />

8.5<br />

EO12SO10 56.3 5.7 5.6 0.3<br />

EO10SO10EO10 51.1 8.7 4.8 2.4<br />

EO137SO18EO137 75.9 7.1 2.1 0.9<br />

a Estimated uncertainty: τ01, τ01 to (1%; E01, E02 to (0.01.<br />

the loss modulus E′′ is close to zero, in particular, for copolymer<br />

EO137SO18EO137. This involves that the exchange processes<br />

between the surface layer <strong>and</strong> bulk solution are relatively<br />

important, <strong>and</strong> as the frequency becomes larger, the system does<br />

not have enough time to reach the equilibrium <strong>and</strong> returns most<br />

<strong>of</strong> the stored energy. Thus, the elasticity becomes larger.<br />

Otherwise, the time is not long enough to modify the interfacial<br />

concentration gradient through different relaxation processes at<br />

high oscillation frequencies. As a result, the dilational moduli<br />

increase with increasing oscillation frequencies. On the other<br />

h<strong>and</strong>, τ01 <strong>and</strong> τ02 values at the a/w interface for copolymers<br />

EO12SO10, EO10SO10EO10, <strong>and</strong> EO137SO18EO137 derived from<br />

the Maxwell modelization are displayed in Table 2. In particular,<br />

copolymer EO12SO10 possesses the shortest relaxation times τ01<br />

<strong>and</strong> τ02, which indicates the fastest adsorption-desorption<br />

exchange <strong>and</strong> a rapid molecular reorganization <strong>of</strong> the polymer<br />

chains adsorbed at the interface probably as a result <strong>of</strong> its lower<br />

molecular weight, diblock architecture, <strong>and</strong> larger hydrophobicity.<br />

At the c/w interface, the pr<strong>of</strong>iles <strong>of</strong> the dilational rheology<br />

curves display similar behavior as at the a/w interface, although<br />

the absolute values <strong>of</strong> E01 <strong>and</strong> E02 are sensibly lower, particularly<br />

the former ones. For this reason, it seems that a viscous fluidlike<br />

behavior is displayed by the three copolymers at the c/w<br />

interface. Moreover, the values obtained for the characteristic<br />

relaxation times τ01 <strong>and</strong> intrinsic elasticity E01 decreased,<br />

whereas τ02 <strong>and</strong> E02 increased if compared to those obtained at<br />

the a/w interface; that is, a reverse behavior is found. This may<br />

point out that the copolymers have been weakly adsorbed at<br />

the c/w interface because copolymer molecules tend to stay at<br />

the chlor<strong>of</strong>orm phase because chlor<strong>of</strong>orm is a good solvent for<br />

both hydrophilic <strong>and</strong> hydrophobic copolymer blocks, as previously<br />

mentioned.<br />

Adsorption Isotherms. Monolayers <strong>of</strong> EO12SO10, EO10-<br />

SO10EO10, <strong>and</strong> EO137SO18EO137 block copolymers were spread<br />

on a Langmuir-Blodgett trough balance, <strong>and</strong> the Π-A isotherms<br />

were obtained (see Figure 3a). In the present study,<br />

copolymer EO137SO18EO137 displays a classical isotherm pattern<br />

divided in four regions. When no pressure is exerted, copolymer<br />

chains should lie on the interface with a flat (“pancake”)<br />

conformation parallel to the surface plane. 46,47 The area occupied<br />

is a function <strong>of</strong> the number <strong>of</strong> SO <strong>and</strong> EO units. Roughly, the<br />

maximum cross-sectional area occupied by an EO unit is<br />

13-16.5 Å 225 <strong>and</strong> that occupied by an SO unit can be taken<br />

similar as that <strong>of</strong> a PS unit (50 Å 2 ). 48 Once hydrated, the areas<br />

<strong>of</strong> EO <strong>and</strong> SO units increase by 8.5 Å 2 (a water molecule). Once<br />

the compression <strong>of</strong> the EO137SO18EO137 monolayer began, the<br />

surface pressure gradually increased until reaching an area under<br />

compression <strong>of</strong> 7350 Å 2 /molecule. This value agrees with the<br />

117


15708 J. Phys. Chem. C, Vol. 114, No. 37, 2010 Juárez et al.<br />

Figure 3. (a) Surface pressure, Π, isotherms for spread monolayers<br />

<strong>of</strong> copolymers (blue line) EO12SO10, (red line) EO10SO10EO10, <strong>and</strong><br />

(black line) EO137SO18EO137. The inset is an enlargement <strong>of</strong> the isotherm<br />

corresconding to copolymers EO12SO10 <strong>and</strong> EO10SO10EO10. (b) Compressibility,<br />

K, against surface pressure <strong>of</strong> spread monolayers <strong>of</strong><br />

copolymers (red line) EO12SO10, (blue line) EO10SO10EO10, <strong>and</strong> (black<br />

line) EO137SO18EO137.<br />

sum <strong>of</strong> the maximum transverse area occupied by all EO <strong>and</strong><br />

SO units <strong>of</strong> copolymer EO137SO18EO137. As the pressure<br />

increased, the hydrophobic SO blocks, initially on the air-water<br />

interface, were lifted away. The surface pressure at which this<br />

phenomenon occurs is usually very low owing to the weakness<br />

<strong>of</strong> the interaction between the aqueous medium <strong>and</strong> the<br />

hydrophobic groups 49 (Figure 3a, region ii). The 4800 Å 2 /<br />

molecule area corresponds to an arrangement <strong>of</strong> the molecules<br />

that leads to an occupied area 35% lower than the original value.<br />

From 6.1 to 9.8 mN/m, another region <strong>of</strong> a lower slope appears<br />

due to a change in the copolymer conformation, which enables<br />

the penetration <strong>of</strong> the hydrophilic EO chains into the subphase:<br />

the copolymer adopts a “mushroom” conformation (Figure 3a,<br />

region iii). In this region, there is a slight increase in the surface<br />

pressure as the surface area is decreased, indicating that a true<br />

first-order transition does not occur. This “pseudoplateau” is<br />

interpreted as a rearrangement <strong>of</strong> the SO coils into “loops”<br />

within the monolayer regime <strong>and</strong> the immersion <strong>of</strong> more EO<br />

units in the aqueous subphase. The end <strong>of</strong> the pseudoplateau<br />

region correponds to an area <strong>of</strong> ∼1800 Å 2 /molecule, which<br />

indicates that the vast majority <strong>of</strong> EO units is in the subphase.<br />

Isotherms <strong>of</strong> PS-PEO copolymers <strong>of</strong> different PEO lengths<br />

showed that the PEO block length largely influences in the<br />

character <strong>of</strong> this transition. 50 Further compression causes a rapid<br />

increase in surface pressure. Block copolymer molecules gradually<br />

become closer, the mobility <strong>of</strong> the blocks becomes restricted<br />

because <strong>of</strong> both space limitations <strong>and</strong> the increase in lateral<br />

interactions, <strong>and</strong> the copolymer molecules reorganize into a<br />

“brush conformation” (Figure 3a, region iv). 51-53 In the subphase,<br />

the EO chains entangle with neighboring copolymer<br />

molecules, whereas at the interface, the SO blocks can form<br />

loops <strong>and</strong> even are partially solubilized in the aqueous EO layer.<br />

If the area is further restricted, both hydrophilic <strong>and</strong> hydrophobic<br />

blocks become stretched (condensed state). 25 The extrapolation<br />

<strong>of</strong> the experimental data above this pressure indicates that the<br />

area occupied per molecule at the condensed state (AL) is<br />

276 Å 2 .<br />

In contrast, copolymers EO12SO10 <strong>and</strong> EO10SO10EO10 exerted<br />

a relatively low resistance to compression at the largest <strong>and</strong><br />

medium areas per molecule <strong>of</strong> their isotherms (see Figure 3a),<br />

as expected from their low molecular weight, shorter SO <strong>and</strong><br />

EO segments, <strong>and</strong> fair aqueous solubility at low concentrations.<br />

The surface pressures <strong>of</strong> EO12SO10 <strong>and</strong> EO10SO10EO10 monolayers<br />

were measurable at significantly smaller areas per<br />

molecule than for the EO137SO18EO137 monolayer. Only two<br />

well-defined regions can be distinguished in their isotherms.<br />

When the area per molecule is large, the surface pressure slowly<br />

increases as the area decreases. Above 1 <strong>and</strong> 2 mN/m for<br />

EO12SO10 <strong>and</strong> EO10SO10EO10, respectively, a very slight pseudoplateau<br />

might be depicted. This behavior arises from the<br />

increasing weight fraction <strong>of</strong> SO blocks at the interface, which<br />

reduces the fractional interfacial area occupied by EO segments.<br />

Above 3 <strong>and</strong> 5 mN/m, respectively, a sharp rise in the surface<br />

pressure occurs. This is consistent with a significant contribution<br />

from the SO domains, showing no plateau or pseudoplateau at<br />

10 mN/m. In fact, the monolayers <strong>of</strong> both block copolymers<br />

can be considered as condensed-like films in comparison with<br />

that obtained for EO137SO18EO137, which behaves as an exp<strong>and</strong>edlike<br />

film. 53 The area occupied per molecule at the condensed<br />

state (AL) is 167 Å 2 for E12S10 <strong>and</strong> 135 Å 2 for EO10SO10EO10,<br />

respectively. These values are in fair agreement with the areas<br />

predicted from the number <strong>of</strong> SO units <strong>of</strong> EO12SO10 <strong>and</strong><br />

EO10SO10EO10, respectively, <strong>and</strong> support the fact that the<br />

interface is occupied only by SO blocks. This is also confirmed<br />

by the fact that, despite that the isotherms <strong>of</strong> the copolymers<br />

clearly differ in their shape at low surface pressures, they merge,<br />

becoming almost superimposable at large surface pressures.<br />

On the other h<strong>and</strong>, the compressibility modulus (K )-1/A<br />

× (dA/dπ)) displays local maxima for every phase transition in<br />

the copolymers’ monolayers. As shown in Figure 3b, several<br />

phase transitions located at ∼6.5, ∼9.5, <strong>and</strong> ∼14.0 m/Nm were<br />

found for copolymer EO137SO18EO137. Those at ∼6.5 <strong>and</strong> ∼9.5<br />

m/Nm are PEO-related phase transitions, which are a consequence<br />

<strong>of</strong> the penetration <strong>of</strong> PEO chains in the subphase, leaving<br />

the air-water interface to form a swollen three-dimensional<br />

structure in the aqueous phase (in particular, at ∼9.5 m/mN).<br />

After the latter peak, the subsequent minimum <strong>and</strong> additional<br />

maximum found at larger surface pressures is a result <strong>of</strong> the<br />

increasing repulsion between SO segments in the top layer <strong>and</strong><br />

PEO segments in the subphase. This facilitates the desorption<br />

<strong>of</strong> SO blocks due to enhanced repulsions within the layer <strong>and</strong><br />

the formation <strong>of</strong> loops <strong>and</strong> tails, which leads to a faster<br />

redistribution between the upper <strong>and</strong> lower regions <strong>of</strong> the layer.<br />

On the other h<strong>and</strong>, only slight maxima at ∼2.0 m/mN <strong>and</strong> a<br />

shoulder at ca. 6.0 m/mN are found for copolymers EO12SO10<br />

<strong>and</strong> EO10SO10EO10. The maximum can be related to a PEO<br />

phase transition, whereas the shoulder at 6.5 m/mN can resemble<br />

118


Amphiphilic PEO-PSO Block Copolymers J. Phys. Chem. C, Vol. 114, No. 37, 2010 15709<br />

Figure 4. Surface pressure isotherms plotted with respect to (a) the<br />

average area per EO repeat unit <strong>and</strong> (b) the average area per SO repeat<br />

unit for copolymers (blue line) EO12SO10, (red line) EO10SO10EO10,<br />

<strong>and</strong> (black line) EO137SO18EO137.<br />

some kind <strong>of</strong> structural rearrangement <strong>of</strong> SO blocks before<br />

complete film condensation. The absence <strong>of</strong> the transition at<br />

9.5 m/Nm for both EO12SO10 <strong>and</strong> EO10SO10EO10 is a consequence<br />

<strong>of</strong> their small PEO block length.<br />

The effect <strong>of</strong> the polymer structure can be easily visualized<br />

by plotting the surface pressure as a function <strong>of</strong> the average<br />

surface area per monomer. Isotherms plotted with the molecular<br />

area normalized with respect to the number <strong>of</strong> EO <strong>and</strong> SO<br />

segments are shown in Figure 4a,b. Figure 4a demonstrates that<br />

all <strong>of</strong> the isotherms <strong>of</strong> the three copolymers fall on a single<br />

curve for surface areas corresponding to the semidilute regime.<br />

This indicates that the isotherm in this regime is mainly<br />

dependent on the EO block length. Beyond the pseudoplateau,<br />

all copolymers appear to be affected by the SO/EO ratio. In the<br />

semidilute regime, the Flory coefficient <strong>of</strong> the system can be<br />

obtained from the des Cloizeaux equation, 54 which describes<br />

the relationship between the surface pressure <strong>and</strong> the surface<br />

concentration (Γ) <strong>of</strong> a polymer by<br />

Π ) CA -y ) KΓ y with y ) dν/(dν - 1) (7)<br />

where C <strong>and</strong> K are proportionality constants, A is the molecular<br />

area, y is the scaling exponent, d is the geometrical dimension,<br />

<strong>and</strong> ν is the Flory coefficient used to express the radius <strong>of</strong><br />

gyration in terms <strong>of</strong> molecular weight (Rg ≈ M ν ) <strong>and</strong> is a<br />

measure <strong>of</strong> the solvent quality. 55 For chains in good solvent<br />

conditions, the exponent ν is theoretically predicted to be 0.75<br />

for a 2D self-avoiding walk <strong>and</strong> 0.6 for a 3D self-avoiding walk.<br />

Consequently, the theoretical exponent for the scaling <strong>of</strong> the<br />

surface pressure with the area per molecule is y ) 3 for the 2D<br />

semidilute regime <strong>and</strong> y ) 2.25 for the 3D semidilute regime.<br />

Plotting the variation <strong>of</strong> the surface pressure Π versus the<br />

inverse <strong>of</strong> the area per molecules 1/A on a double-logarithmic<br />

scale for the three different copolymers in the intermediate<br />

region between the dilute regime <strong>and</strong> the plateau, we have<br />

obtained y exponents <strong>of</strong> 2.28, 2.37, <strong>and</strong> 2.54 for EO12SO10,<br />

EO10SO10EO10, <strong>and</strong> EO137SO18EO137, respectively. This indicates<br />

that the chain conformation is almost a 3D interpenetrated layer<br />

for copolymers EO12SO10 <strong>and</strong> EO10SO10EO10 <strong>and</strong> a intermediate<br />

between a purely 2D entangled layer <strong>and</strong> a 3D interpenetrated<br />

layer for copolymer EO137SO18EO137, as also observed for<br />

different PSO-EO block copolymers. 56 From Figure 4b, the<br />

isotherms superimposed in the concentrated regime for the three<br />

block copolymers, which denotes that the isotherm shape is now<br />

determined by the SO chain length <strong>and</strong> SO/EO molar ratio.<br />

Summarizing, if the SO block is short compared with the<br />

EO block, as in the case <strong>of</strong> copolymer EO137SO18EO137, its role<br />

is limited to anchoring the EO chains at the air-water interface<br />

<strong>and</strong> it has no influence on the surface pressure isotherm or on<br />

the segment pr<strong>of</strong>ile <strong>of</strong> the EO chains normal to the interface.<br />

But, as the SO block is repulsive to the EO block, as in the<br />

case <strong>of</strong> EO12SO10 <strong>and</strong> EO10SO10EO10, this should modify<br />

the attractive nature <strong>of</strong> the interface for the EO <strong>and</strong>, therefore,<br />

the expected concentration pr<strong>of</strong>ile for the EO chains, as also<br />

occurred for PS-EO <strong>and</strong> EO-PS-EO copolymers 34,47,49,57,58 <strong>and</strong><br />

Pluronics 26 <strong>and</strong> Tetronics (four star-shaped ethylene oxidepropylene<br />

oxide block copolymers). 28 The balance between these<br />

opposite interactions will depend on the copolymer surface<br />

density, the SO block length, <strong>and</strong> the temperature.<br />

AFM. Besides the isotherm data, detailed insight into the<br />

organization <strong>and</strong> morphology <strong>of</strong> the Langmuir-Blodgett monolayers<br />

should rely on the images <strong>of</strong> their corresponding films<br />

by AFM. We compare the morphologies <strong>of</strong> the three samples<br />

deposited at different pressures. Figure 5 shows AFM images<br />

<strong>of</strong> the block copolymer films on freshly cleaved mica obtained<br />

at 5 <strong>and</strong> 11 mN/m surface pressures. Aggregation was observed<br />

for all the copolymers investigated. The surface micelle assembly<br />

can take place (i) during the solvent evaporation step<br />

or (ii) during the LB tranfer. Nevertheless, dynamic lightscattering<br />

measurements (DLS, data not shown) <strong>of</strong> the three<br />

copolymers in chlor<strong>of</strong>orm did not show any evidence <strong>of</strong> micelle<br />

formation, which implies that the formation <strong>of</strong> EO-SO surface<br />

features in LB films are the result <strong>of</strong> spontaneous copolymer<br />

aggregation at the air-water interface rather than a transfer <strong>of</strong><br />

micelles formed in the spreading solution. 59<br />

At a surface pressure <strong>of</strong> 5 mN/m, the AFM images <strong>of</strong> the<br />

block copolymers EO12SO10, EO10SO10EO10, <strong>and</strong> EO137-<br />

SO18EO137 show the presence <strong>of</strong> spherical aggregates with<br />

diameters <strong>of</strong> 6.7 ( 0.4, 7.9 ( 0.2, <strong>and</strong> 9.1 ( 0.3 nm, respectively<br />

(Figure 5a,c,e; see Figure S1 in the Supporting Information for<br />

size distributions). The average height <strong>of</strong> the aggregates obtained<br />

at this surface pressure is 2.03 ( 0.01, 1.25 ( 0.01, <strong>and</strong> 1.52<br />

( 0.01 nm for EO12SO10, EO10SO10EO10, <strong>and</strong> EO137SO18EO137,<br />

respectively, so they can be identified as dots or circular micelles<br />

(see Figure 6 as an example <strong>and</strong> Figure S2 in the Supporting<br />

Informationfor height distributions). The difference in dot<br />

heights for the three copolymers is a consequence <strong>of</strong> the<br />

119


15710 J. Phys. Chem. C, Vol. 114, No. 37, 2010 Juárez et al.<br />

Figure 5. Typical AFM images <strong>of</strong> Langmuir-Blodgett films <strong>of</strong> the<br />

copolymers EO12SO10 (a, b), EO137SO18EO137 (c, d), <strong>and</strong> EO10SO10EO10<br />

(e, f) at transfer surface pressures <strong>of</strong> 5 <strong>and</strong> 11 mN/m, respectively.<br />

Arrows denote some elongated polymer structures.<br />

copolymer hydrophobicity <strong>and</strong> SO/EO ratio: Assuming that the<br />

length <strong>of</strong> an SO unit is 0.18 nm per chain unit, 60 the average<br />

length <strong>of</strong> the fully stretched SO10 <strong>and</strong> SO18 blocks would be<br />

1.8 <strong>and</strong> 3.3 nm. As the central block is looped in the aggregate<br />

core in the case <strong>of</strong> the triblock copolymers, the effective length<br />

would be 0.9 <strong>and</strong> 1.7 nm, respectively. On comparison with<br />

the experimental values, the SO blocks <strong>of</strong> copolymers EO12SO10<br />

<strong>and</strong> EO10SO10EO10 display a more extended perpendicular<br />

configuration on the water surface (“brush” state) than for<br />

copolymer EO137SO18EO137, for which a “mushroom” conformation<br />

is predominant, confirming the surface-pressure isotherm<br />

data.<br />

When the transfer surface pressure rises to 11 mN/m, the<br />

height <strong>of</strong> the EO12SO10 layer largely increases up to 5.2 ( 0.01<br />

nm, which suggests the formation <strong>of</strong> multilayers at the a/w<br />

interface. Formation <strong>of</strong> large aggregates is also observed as a<br />

consequence <strong>of</strong> micelle association. Part <strong>of</strong> these supraaggregates<br />

display an elongated morphology, which seems to<br />

be composed <strong>of</strong> a string <strong>of</strong> beads, the beads being circular<br />

micelles in contact each other. Similar necklace-network strctures<br />

have been observed, for example, in Langmuir-Blodgett<br />

films <strong>of</strong> mixed PS <strong>and</strong> PS-P2VP (P2VP, poly(2-vinylpyridine)).<br />

61 This large-scale aggregation pattern is observed,<br />

suggesting the possibility <strong>of</strong> a dewetting process, although we<br />

cannot neglect the possible influence <strong>of</strong> particle aggregation<br />

upon film drying. In the case <strong>of</strong> copolymers EO10SO10EO10 <strong>and</strong><br />

EO137SO18EO137, the increase in film thickness increase upon<br />

transfer at 11 mN/m is less pronounced, with mean height values<br />

<strong>of</strong> 1.95 ( 0.01 <strong>and</strong> 3.90 ( 0.02 nm, respectively. These values<br />

point out that the SO blocks <strong>of</strong> both copolymers are in a full<br />

extended perpendicular configuration (brush state) at this transfer<br />

surface pressure. The average size <strong>of</strong> the aggregates is 7.2 (<br />

0.1, 5.6 ( 0.3, <strong>and</strong> 7.3 ( 0.1 nm for EO12SO10, EO10SO10EO10,<br />

<strong>and</strong> EO137SO18EO137, respectively, at this transfer pressure.<br />

Comparing these values with those measured at a surface<br />

pressure <strong>of</strong> 5 mN/m, we can assume that a compaction <strong>of</strong> the<br />

aggregates formed by copolymers EO10SO10EO10 <strong>and</strong><br />

EO137SO18EO137 has occurred in order to accommodate all<br />

copolymer chains at the interface. Moreover, formation <strong>of</strong><br />

elongated micelles is also observed at 11 mN/m for the former<br />

copolymers (see Figure 5d,f). In fact, elongated micelles in bulk<br />

solution have been previously found for copolymer<br />

EO10SO10EO10, whereas for EO137SO18EO137, only spherical<br />

micelles were detected. 32<br />

In contrast to what was found by Deveraux et al. 60 <strong>and</strong> M<strong>of</strong>fitt<br />

et al. 33 for PS-EO block copolymers, the deposition pressure<br />

influences the structure <strong>of</strong> the resulting copolymer surface<br />

aggregates. Furthermore, PS-EO copolymers show additional<br />

surface structures other than dots (elongated structures, lamellae...)<br />

when the EO weight contents are lower than 12%; 33<br />

however, for SO-based copolymers, elongated structures are<br />

Figure 6. Determination <strong>of</strong> micelle size for copolymer EO137SO18EO137 at a surface pressure <strong>of</strong> 11 mN/m through sectional analysis. X represents<br />

the aggregate diameter <strong>and</strong> Z the aggregate height.<br />

120


Amphiphilic PEO-PSO Block Copolymers J. Phys. Chem. C, Vol. 114, No. 37, 2010 15711<br />

TABLE 3: Aggregate Number, Nagg, for Surface Micelles <strong>of</strong><br />

Copolymers EO12SO10, EO10SO10EO10, <strong>and</strong> EO137SO18EO137<br />

Nagg<br />

5 mN/m 11 mN/m<br />

EO12SO10 40 144<br />

EO10SO10EO10 29 68<br />

EO137SO18EO137 10 36<br />

already observed at an EO content <strong>of</strong> 27% for copolymer<br />

EO12SO10. The lower glass transition temperature <strong>of</strong> SO (∼40<br />

°C) compared with that <strong>of</strong> a PS (∼100 °C), which is thought to<br />

be associated with plasticization <strong>of</strong> water molecules associated<br />

with the ether oxygen <strong>of</strong> the SO units, 12 might involve a larger<br />

mobility <strong>of</strong> SO blocks in the aggregates, thus enabling a certain<br />

restructurization <strong>of</strong> the aggregates consisting <strong>of</strong> lateral interactions<br />

between SO segments <strong>of</strong> adjacent circular aggregates.<br />

The aggregation number (Nagg) <strong>of</strong> the copolymer aggregates<br />

was evaluated from the AFM images in conjuction with the<br />

Π-A isotherms. Nagg was determined by dividing the number<br />

<strong>of</strong> micelles per unit area by the molecular area from the Π-A<br />

isotherms at the corresponding transfer pressure. The height<br />

pr<strong>of</strong>ile <strong>of</strong> the micelles was generally found to be hemispheric<br />

with the condition used to scan the samples. From Table 3, it<br />

can be seen that Nagg increases with the increase <strong>of</strong> the SO<br />

weight fraction at a certain deposition pressure, which can be a<br />

consequence <strong>of</strong> stronger attractive interactions between the SO<br />

blocks to avoid contact with the solvent. Values <strong>of</strong> Nagg are in<br />

fair agreement with those obtained for micelles in solution. 32<br />

The lower aggregation number <strong>of</strong> copolymer EO137SO18EO137<br />

may arise from the steric repulsion <strong>of</strong> EO chains at the interface<br />

due to the longer PEO blocks if compared to the other block<br />

copolymer. Moreover, Nagg for the corresponding samples tends<br />

to increase with the increase <strong>of</strong> the deposition pressure because<br />

further compression will provide enough energy to overcome<br />

repulsion between EO segments <strong>and</strong> form larger aggregates.<br />

Conclusions<br />

In summary, the spontaneous adsorption <strong>of</strong> block copolymers<br />

EO12SO10,EO10SO10EO10, <strong>and</strong> EO137SO18EO137 at the air-water<br />

interface is observed to be slowed down when the hydrophobicity<br />

<strong>of</strong> the molecule is increased. In contrast, at the chlor<strong>of</strong>ormwater<br />

interface, no measurable effect is observed for copolymer<br />

EO12SO10 due to the slow diffusion <strong>of</strong> its chains to the interface<br />

in combination with a low bulk concentration, the slow diffusion<br />

associated with the fact that chlor<strong>of</strong>orm is a good solvent for<br />

both EO <strong>and</strong> SO blocks. In addition, the adsorption layers at<br />

the a/w interface manifest obvious solid-like properties in the<br />

whole accessible frequency range with E′ >E′′, following the<br />

sequence EO12SO10 > EO10SO10EO10 > EO137SO18EO137, <strong>and</strong><br />

with a low significance <strong>of</strong> dissipative processes. In contrast, a<br />

viscous fluid-like behavior is displayed by the three copolymers<br />

at the c/w interface. Copolymer EO137SO18EO137 displays an<br />

adsorption isotherm with the four classical regions representing<br />

pancake, mushroom, brush, <strong>and</strong> condensed states, with the<br />

presence <strong>of</strong> a pseudoplateau attributed to the pancake-to-brush<br />

transition as EO chains submerge into the aqueous subphase.<br />

On the other h<strong>and</strong>, for EO12SO10 <strong>and</strong> EO10SO10EO10 copolymers,<br />

only two regions are observed in their adsorption isotherms as<br />

a consequence <strong>of</strong> their low molecular weight, short SO <strong>and</strong> EO<br />

block lengths, <strong>and</strong> much larger SO/EO ratio. The latter involves<br />

the disappearance <strong>of</strong> the pseudoplateau region due to the<br />

decrease in the fractional interfacial area occupied by EO<br />

segments. Finally, surface circular micelles are observed by<br />

AFM pictures at two surface transfer pressures. A decrease in<br />

micelle size <strong>and</strong> an increase in monolayer thickness are observed<br />

with an increase in transfer pressure. In addition, at the largest<br />

transfer pressure, elongated micelles are observed. In the case<br />

<strong>of</strong> copolymer EO12SO10, association <strong>of</strong> micelles is also observed.<br />

Aggregation numbers derived from AFM images increase with<br />

the increase <strong>of</strong> the SO weight fraction at a certain deposition<br />

pressure, which can be a consequence <strong>of</strong> stronger attractive<br />

interactions between the SO blocks to avoid contact with the<br />

solvent. The lower aggregation number <strong>of</strong> copolymer<br />

EO137SO18EO137 may arise from the steric repulsion <strong>of</strong> EO<br />

chains at the interface due to the longer PEO blocks if compared<br />

to the other block copolymer.<br />

Acknowledgment. The authors thank the Ministerio de<br />

Ciencia e Innovación (MICINN) for research projects MAT<br />

2007-6107 <strong>and</strong> Xunta de Galicia for additional financial support<br />

under project INCITE09206020PR. S.G.-L. thanks the MICINN<br />

for her FPI scholarship. The authors also thank Pr<strong>of</strong>. D. Attwood<br />

<strong>and</strong> C. Booth for the generous gift <strong>of</strong> the copolymers.<br />

Supporting Information Available: Size <strong>and</strong> height distributions<br />

<strong>of</strong> aggregates formed in Langmuir-Blodgett films<br />

<strong>of</strong> copolymers at surface pressures <strong>of</strong> 5 <strong>and</strong> 11 mN/M. This<br />

material is available free <strong>of</strong> charge via the Internet at http://<br />

pubs.acs.org.<br />

References <strong>and</strong> Notes<br />

(1) Edens, M. W. Nonionic Surfactants, Polyoxyalkylene Block Copolymers;<br />

Marcel Dekker: New York, 1996.<br />

(2) Riess, G.; Cheymol, A.; Hoerner, P.; Krikorian, R. AdV. Colloid<br />

Interface Sci. 2004, 108-109, 43–48.<br />

(3) Tao, Y.; Ma, B.; Segalman, R. A. Macromolecules 2008, 41, 7152–<br />

7159.<br />

(4) Boudier, A.; Aubert-Pouëssel, A.; Gérardin, C.; Devoisselle, J. M.;<br />

Bégu, S. Int. J. Pharm. 2009, 379, 212–217.<br />

(5) Gilcreest, V. P.; Dawson, K. A.; Gorelov, A. V. J. Phys. Chem. B<br />

2006, 110, 21903–21910.<br />

(6) George, P. A.; Donose, B. C.; Cooper-White, J. J. Biomaterials<br />

2009, 30, 2449–2456.<br />

(7) Sedev, R.; Jachimska, B.; Khristov, K.; Malysa, K.; Exrowa, D. J.<br />

Dispersion Sci. Technol. 1999, 20, 1759–1776.<br />

(8) Alex<strong>and</strong>ridis, P.; Lindman, B. Amphiphilic Block Copolymers: <strong>Self</strong>-<br />

<strong>Assembly</strong> <strong>and</strong> Applications; Elsevier Science: Amsterdam, 2000.<br />

(9) Park, S.; Wang, J. W.; Kim, B.; Russell, T. P. Nano Lett. 2008, 8,<br />

1667–1672.<br />

(10) Park, S.; Kim, B.; Wang, J. W.; Russell, T. P. AdV. Mater. 2008,<br />

20, 681–685.<br />

(11) Hillmyer, H. A. AdV. Polym. Sci. 2005, 190, 137.<br />

(12) Crothers, M.; Attwood, D.; Collet, J. H.; Yang, Z.; Booth, C.;<br />

Taboada, P.; Mosquera, V.; Ricardo, N. P. S.; Martini, L. G. A. Langmuir<br />

2002, 18, 8685–8691.<br />

(13) Yang, Z. M.; Crothers, M.; Ricardo, N. P. S.; Chaibundit, C.;<br />

Taboada, P.; Mosquera, V.; Kelarakis, A.; Havredaki, V.; Martini, L.;<br />

Valder, C.; Collett, J. H.; Attwood, D.; Heatley, F.; Booth, C. Langmuir<br />

2003, 19, 943–950.<br />

(14) Booth, C.; Attwood, D.; Price, C. Phys. Chem. Chem. Phys. 2006,<br />

8, 3612–3622.<br />

(15) Riess, G. Prog. Polym. Sci. 2003, 28, 1107–1170.<br />

(16) Booth, C.; Attwood, D. Macromol. Rapid Commun. 2000, 21, 501–<br />

527.<br />

(17) Gangaly, R.; Awal, V. K.; Hassan, P. A. J. Colloid Interface Sci.<br />

2007, 325, 693–700.<br />

(18) Battaglia, G.; Ryan, A. J. J. Am. Chem. Soc. 2005, 127, 8757–<br />

8764.<br />

(19) Bryskhe, K.; Jansson, J.; Topgaard, D.; Schillén., K.; Olsson, U.<br />

J. Phys. Chem. B 2004, 108, 9710–9719.<br />

(20) Chaibundit, C.; Sumanatrakool, P.; Chinchew, S.; Kanatharana, P.;<br />

Tattershall, C. E.; Booth, C.; Yuan, Y.-F. J. Colloid Interface Sci. 2005,<br />

283, 544–554.<br />

(21) Barbosa, S.; Cheema, M. A.; Taboada, P.; Mosquera, V. J. Phys.<br />

Chem. B 2007, 111, 10920–10928.<br />

(22) Vaughn, T. G.; suter, H. R.; Lunsted, L. G.; Kramer, M. G. J. Am.<br />

Oil Chem. Soc. 1951, 28, 294.<br />

(23) Nace, V. M. J. Am. Oil Chem. Soc. 1996, 73, 1–8.<br />

121


15712 J. Phys. Chem. C, Vol. 114, No. 37, 2010 Juárez et al.<br />

(24) Fleute-Schlachter, I. Goldschmidt Industrial Specialties, Essen,<br />

Germany. Private communication.<br />

(25) O′Connor, S. M.; Gehrke, S. H.; Retzinger, G. S. Langmuir 1999,<br />

15, 2580–2585.<br />

(26) Rippner Blomquist, B.; Wärnheim, T.; Claesson, P. M. Langmuir<br />

2005, 21, 6373–6384.<br />

(27) González-López, J.; S<strong>and</strong>ez-Macho, I.; Concheiro, A.; Alvarez-<br />

Lorenzo, C. J. Phys. Chem. C 2010, 114, 1181–1189.<br />

(28) González-López, J.; Alvarez-Lorenzo, C.; Taboada, P.; Sosnik, A.;<br />

S<strong>and</strong>ez-Macho, I.; Concheiro, A. Langmuir 2008, 24, 10688–10697.<br />

(29) Hodges, C. S.; Neville, F.; Konovalov, O.; Hammond, R. B.;<br />

Gidalevitz, D.; Hamley, I. W. Langmuir 2006, 22, 8821–8825.<br />

(30) Ricardo, N. M. P. S.; Chaibundit, C.; Yang, Z.; Attwood, D.; Booth,<br />

C. Langmuir 2006, 22, 1301–1306.<br />

(31) Pinho, M. E. N.; Costa, F. M. L. L.; Filho, F. B. S.; Ricardo,<br />

N. M. P. S.; Yeates, S. G.; Attwood, D.; Booth, C. Int. J. Pharm. 2007,<br />

328, 95–98.<br />

(32) Juárez, J.; Taboada, P.; Valdez, M. A.; Mosquera, V. Langmuir<br />

2008, 24, 7107–7116.<br />

(33) Cheyne, R. B.; M<strong>of</strong>fitt, M. G. Langmuir 2005, 21, 5453–5460.<br />

(34) Deschênes, L.; Bousmina, M.; Ritcey, A. M. Langmuir 2008, 24,<br />

3699–3708.<br />

(35) Crothers, M.; Zhou, Z.; Ricardo, N. P. S.; Yang, Z.; Taboada, P.;<br />

Chaibundit, C.; Attwood, D.; Booth, C. Int. J. Pharm. 2005, 293, 91–100.<br />

(36) Elsabahy, M.; Perron, M.-È.; Bertr<strong>and</strong>, N.; Yu, G.-e.; Leroux, J.-<br />

C. Biomacromolecules 2007, 8, 2250–2257.<br />

(37) Babak, V. G.; Boury, F. Colloids Surf., A 2004, 243, 33–42.<br />

(38) Babak, V. G.; Baros, F.; Boury, F.; Desbrières, J. Colloids Surf., B<br />

2008, 65, 43–49.<br />

(39) Ward, A. F. H.; Tordai, L. J. Chem. Phys. 1946, 14, 453–458.<br />

(40) Hansen, F. K.; Myrvold, R. J. Colloid Interface Sci. 1995, 176,<br />

498.<br />

(41) Myrvold, R.; Hansen, F. K.; Balinov, B. Colloids Surf., A 1996,<br />

18, 453.<br />

(42) Muñoz, M. G.; Monroy, F.; Ortega, F.; Rubio, R. G.; Langevin,<br />

D. Langmuir 2000, 16, 1094–1101.<br />

(43) Joos, P.; Fainerman, V. B. Dynamic Surface Phenomena; VSP:<br />

Utrecht, The Netherl<strong>and</strong>s, 1999.<br />

(44) Kopperud, H. B. M.; Hansen, F. K. Macromolecules 2001, 34,<br />

5635–5643.<br />

(45) Leiva, A.; Farias, A.; Gargallo, L.; Radic, D. Eur. Polym. J. 2008,<br />

44, 2589–2598.<br />

(46) Haefele, T.; Kita-Tokarczyk, K.; Meier, W. Langmuir 2006, 22,<br />

1164–1172.<br />

(47) Kiss, E.; Keszthelyi, T.; Kormany, G.; Hakkel, O. Macromolecules<br />

2006, 39, 9375–9384.<br />

(48) Elbolk, T. A.; Detellier, C. J. Phys. Chem. Solids. 2006, 67, 950–<br />

955.<br />

(49) Hann, R. A. Molecular Structure <strong>and</strong> Monolayer Properties. In<br />

Langmuir-Blodgett Films; Roberts, G., Ed.; Plenum Press: New York, 1990;<br />

pp 18-23.<br />

(50) Nivaggioli, T.; Tsao, B.; Alex<strong>and</strong>ridis, P.; Hatton, A. T. Langmuir<br />

1995, 11, 119–126.<br />

(51) Chen, C.; Even, M. A.; Chen, Z. Macromolecules 2003, 36, 4478–<br />

4484.<br />

(52) Chang, L.-C.; Lin, C. Y.; Kuo, M.-W.; Gau, C.-S. J. Colloid<br />

Interface Sci. 2005, 285, 640–652.<br />

(53) Chang, L. C.; Chang, Y. Y.; Gau, C. S. J. Colloid Interface Sci.<br />

2008, 322, 263–273.<br />

(54) des Cloizeaux, J. J. Phys. (Paris) 1975, 36, 281–291.<br />

(55) Vilanove, R.; Rondelez, F. Phys. ReV. Lett. 1980, 45, 1502.<br />

(56) Faure, M. C.; Bassereau, P.; Lee, L. T.; Menelle, A.; Lheveder, C.<br />

Macromolecules 1999, 32, 8538–8550.<br />

(57) Goncalves da Silva, A. M.; Simoes Gamboa, A. L.; Matinho, J. M.<br />

Langmuir 1998, 14, 5327–5330.<br />

(58) Baker, S. M.; Leach, K. A.; Devereaux, C. E.; Gragson, D. E.<br />

Macromolecules 2000, 33, 5432–5436.<br />

(59) Cox, J. K.; Yu, K.; Constantine, B.; Eisenberg, A.; Lennox, R. B.<br />

Langmuir 1999, 15, 7714–7718.<br />

(60) Devereaux, C. A.; Baker, S. M. Macromolecules 2002, 35, 1921–<br />

1927.<br />

(61) Wen, G.; Chung, B.; Chang, T. Macromol. Rapid Commun. 2008,<br />

29, 1248–1253.<br />

JP1049777<br />

122


4.3.- Appended papers.<br />

In the present work, the solution behaviour <strong>and</strong> surface adsorption properties <strong>of</strong> three different<br />

block copolymer (E12S10, E10S10E10 <strong>and</strong> E137S18E137), have been studied. These block copolymers<br />

display different volume-to-length ratios <strong>of</strong> the hydrophobic block <strong>and</strong> <strong>of</strong> relative lengths <strong>of</strong> the<br />

poly(oxyethylene) block ; <strong>and</strong> also polymer architecture may have an important role in controlling<br />

the morphology <strong>of</strong> polymeric aggregates <strong>and</strong> surface adsorption behaviour <strong>of</strong> these copolymers.<br />

To known the influence <strong>of</strong> such parameters, we studied the self-assembly properties <strong>and</strong> the<br />

different structural aggregates <strong>of</strong> the three copolymers in aqueous solution.<br />

For copolymer E12S10, coexistence <strong>of</strong> spherical micelles with vesicular structures is detected. To the<br />

best <strong>of</strong> our knowledge, the spontaneous formation <strong>of</strong> vesicles by poly(oxystyrene)-<br />

poly(oxyethylene) block copolymers has been first reported here. On the other h<strong>and</strong>, formation <strong>of</strong><br />

elongated micelles upon self-assembly <strong>of</strong> copolymer E10S10E10 in dilute solution is elucidated from<br />

light scattering <strong>and</strong> transmission electron techniques. In the case <strong>of</strong> E137S18E137, typical spherical<br />

micelles are observed, as expected. The micellization properties <strong>of</strong> the latter copolymer had been<br />

already partially studied previously (89)(90), <strong>and</strong> the present study completes this gap.<br />

On the other h<strong>and</strong>, fundamental studies <strong>of</strong> polystyrene oxide-based block copolymers performed<br />

at either the air-water or the water-organic solvent interface can provide valuable information on<br />

their interfacial phase behaviour to further guide their possible use in different biomedical <strong>and</strong><br />

coating applications. To fill this gap, in the present work, we studied the surface behaviour <strong>and</strong><br />

surface properties <strong>of</strong> these three copolymers, E12S10, E10S10E10 <strong>and</strong> E137S18E137, by different<br />

techniques.<br />

To do this work, different techniques such as surface tension, Langmuir-Blodgett through, drop<br />

tensiometry, light scattering, transmission <strong>and</strong> scanning electron microscopies (TEM <strong>and</strong> SEM),<br />

atomic force microscopy (AFM), polarized optical microscopy (POM) <strong>and</strong> rheometry were used.<br />

123


4.3.1.- Relevant aspects<br />

I. For E12S10 dilute solutions, the coexistence <strong>of</strong> spherical micelles with vesicular structures<br />

has been observed. In addition, the spontaneous formation <strong>of</strong> vesicles by<br />

poly(oxystyrene)-poly(oxyethylene) block copolymers is reported for the first time, as<br />

confirmed by light scattering, polarized light microscopy, <strong>and</strong> transmission <strong>and</strong> cryo-<br />

scanning electron microscopy data.<br />

II. In dilute solution, the self-assembly <strong>of</strong> copolymer E10S10E10 leads to the formation <strong>of</strong><br />

elongated micelles as supported by light scattering <strong>and</strong> transmission electron techniques.<br />

III. In the case <strong>of</strong> copolymer E137S18E137, typical spherical micelles are observed. Tube inversion<br />

<strong>and</strong> rheological measurements were used to define the sol- s<strong>of</strong>t- <strong>and</strong> gel-hard boundaries<br />

<strong>of</strong> this copolymer.<br />

IV. E12S10 <strong>and</strong> E10S10E10 copolymers did not form gels in the concentration range analysed.<br />

However, only certain concentrations <strong>of</strong> copolymer E10S10E10 were analysed by rheometry.<br />

From experimental data, an upturn in the low-frequency range <strong>of</strong> the stress moduli was<br />

observed, denoting the existence <strong>of</strong> an emerging slow process, which was assigned to the<br />

formation <strong>of</strong> an elastic network.<br />

V. Block copolymers E12S10, E10S10E10 <strong>and</strong> E137S18E137 showed spontaneous adsorption at the<br />

air-water interface, which slowed down when the hydrophobicity <strong>of</strong> the molecule was<br />

increased. In contrast, at the chlor<strong>of</strong>orm/water interface no measurable effect is observed<br />

for copolymer E12S10 due to the slow diffusion <strong>of</strong> its chains to the interface in combination<br />

with a low bulk concentration; this slow diffusion is associated with the fact that<br />

chlor<strong>of</strong>orm is a good solvent for both E <strong>and</strong> S blocks.<br />

VI. Copolymer E137S18E137 displays an adsorption isotherm with the four classical regions<br />

representing the pancake, mushroom, brush <strong>and</strong> condensed states; the presence <strong>of</strong> a<br />

pseudo-plateau is attributed to the pancake-to-brush transition as E chains submerge into<br />

the aqueous sub-phase. On the other h<strong>and</strong>, for E12S10 <strong>and</strong> E10S10E10 copolymers only two<br />

regions are observed in their adsorption isotherms as a consequence <strong>of</strong> their low<br />

molecular weights, short S <strong>and</strong> E block lengths, <strong>and</strong> much larger S/E ratio. This involves the<br />

disappearance <strong>of</strong> the pseudo-plateau region due to the decrease in the fractional<br />

interfacial area occupied by EO segments.<br />

VII. According to AFM images, circular micelles are observed on Langmuir-Blodgett films <strong>of</strong> the<br />

copolymers obtained at two surface transfer pressures. A decrease in micelle size <strong>and</strong> an<br />

124


increase in monolayer thickness are observed with increases in transfer pressure.<br />

Aggregation numbers derived from AFM images increase with the increase <strong>of</strong> the S weight<br />

fraction at a certain deposition pressure, which can be a consequence <strong>of</strong> stronger<br />

attractive interactions between the S blocks to avoid contact with the solvent.<br />

125


Contents<br />

5.1<br />

5.2<br />

5.3<br />

5.4<br />

5.5<br />

5.1.1<br />

5.1.2<br />

5.1.3<br />

5.1.4<br />

5.2.1<br />

5.2.2<br />

5.2.3<br />

5.2.4<br />

5.3.1<br />

5.3.2<br />

5.5.1<br />

Protein structure<br />

Primary structure <strong>of</strong> the proteins<br />

Secondary structure <strong>of</strong> the proteins<br />

Tertiary structure <strong>of</strong> the proteins<br />

Quaternary structure <strong>of</strong> the proteins<br />

Protein aggregation<br />

Unfolded state<br />

Intermediate <strong>and</strong> transition state assembles<br />

Native state<br />

Energy l<strong>and</strong>scape: protein aggregation<br />

Amyloid fibrils<br />

5.1.- Protein structure<br />

Amyloid fibrils hallmarks<br />

Kinetic fibril formation in vitro<br />

Human serum albumin<br />

Amyloid fibrils as biomaterials<br />

Hen egg-white lysozyme<br />

CHAPTER 5<br />

Proteins<br />

From a chemical point <strong>of</strong> view, proteins are linear heteropolymers whose backbones are<br />

usually composed <strong>of</strong> almost 20 different monomers, as opposed to most synthetic polymers<br />

which are composed <strong>of</strong> one or few monomers. Whilst polymers have a large extended<br />

organization in aqueous solution, proteins fold as relatively small compacted structures. The<br />

spatial arrangement <strong>of</strong> atoms in a protein is called conformation. The possible conformations<br />

<strong>of</strong> a protein include any structural state which can achieve without breaking covalent bonds. A<br />

change in conformation could occur, for example, by rotation about single bonds. The<br />

conformations existing under a given set <strong>of</strong> conditions are usually those most<br />

thermodynamically stable, i.e. having the lowest Gibbs energy.<br />

127<br />

128<br />

129<br />

131<br />

132<br />

132<br />

133<br />

133<br />

134<br />

135<br />

135<br />

135<br />

138<br />

139<br />

141<br />

142<br />

127


Four structural levels are frequently considered in protein architecture. The primary structure<br />

corresponds to the amino acid sequence. The secondary structure refers to the spatial<br />

arrangement <strong>of</strong> amino acid residues that are nearby in the sequence (for example, α-helix <strong>and</strong><br />

β-sheet are elements <strong>of</strong> secondary structure). Tertiary structure refers to the spatial<br />

arrangement <strong>of</strong> amino acid residues that are far apart in the sequence <strong>and</strong> to the pattern <strong>of</strong><br />

disulfide bonds. Finally, proteins containing more than one polypeptide chain exhibit a fourth<br />

level <strong>of</strong> structural organization (each polypeptide chain in such a protein is called a subunit). In<br />

this regard, quaternary structure refers to the spatial arrangement <strong>of</strong> subunits.<br />

5.1.1.- Primary structure <strong>of</strong> the proteins<br />

Proteins are linear polymers formed by the linkage <strong>of</strong> the carboxylic group <strong>of</strong> one amino acid<br />

to the amino group <strong>of</strong> another amino acid generating a peptide bond. The formation <strong>of</strong> a<br />

dipeptide from two amino acids is accompanied by the loss <strong>of</strong> a water molecule. A series <strong>of</strong><br />

amino acids joined by peptide bonds form a polypeptide chain, <strong>and</strong> each amino acid unit in a<br />

polypeptide is called a residue. A polypeptide chain possesses polarity because its ends are<br />

different, with an amino group at one end <strong>and</strong> a carboxylic group at the other. By convention,<br />

the amino end is taken to be the start <strong>of</strong> the polypeptide chain, so the sequence <strong>of</strong> amino<br />

acids in a polypeptide chain is written starting with the amino-terminal residue (4).<br />

Figure 5.1 Polypeptide chain: structural constant backbone (inside the dotted squared box)<br />

<strong>and</strong> lateral side lateral chains, denoted by the letter R.<br />

As mentioned before, proteins are built by a sequence <strong>of</strong> amino acids which are chemically<br />

connected by covalent bonds (peptide bond). A polypeptide chain is constituted <strong>of</strong> two parts: a<br />

main chain or backbone <strong>and</strong> a complementary one constituted by side lateral chains<br />

characteristic <strong>of</strong> each amino acid (Figure 5.1). The backbones possess a great ability to<br />

establish hydrogen bonds: each amino acid has a carboxylic group (-C=O), that is an acceptor<br />

<strong>of</strong> hydrogen bonds (except proline), <strong>and</strong> an amino group (-NH) which it is a donator <strong>of</strong> the<br />

hydrogen bonds. These functional groups mutually interact to provide stability to the<br />

biomacromolecule structure. Naturally, proteins are constituted by 20 different types <strong>of</strong> amino<br />

128


acids whose lateral side chains change in size, form, charge, hydrophobic character <strong>and</strong><br />

chemical reactivity.<br />

5.1.2.- Secondary structure <strong>of</strong> the proteins.<br />

The secondary structure <strong>of</strong> proteins depends on four factors: i) bond length <strong>and</strong> angles <strong>of</strong><br />

peptidic bonds; ii) the coplanar arrangement <strong>of</strong> the substituted atoms in amide groups; iii) the<br />

hydrogen bonding between functional –NH <strong>and</strong> C=O groups in order to maintain structural<br />

stability; iv) the distance <strong>of</strong> the hydrogen bonds that can be formed (3). For instance, segments<br />

<strong>of</strong> polypeptide chains are in a coiled conformation due to intramolecular interaction, i.e. the<br />

atoms involved in amide groups must remain coplanar, <strong>and</strong> for maximal stability, each N-H<br />

group must be hydrogen bonded to a –CO <strong>and</strong> each -CO to a –NH. Under ideal conditions,<br />

functional groups (carboxylic <strong>and</strong> amine groups) <strong>of</strong> amino acids can form hydrogen bonds, with<br />

an energy <strong>of</strong> 5 kcal/mol. The polypeptide chains tend to adopt up to three different<br />

configurations that enable the formation <strong>of</strong> a maximum number <strong>of</strong> hydrogen bonds: alpha<br />

helix, beta sheet, turns or loops (Figure 5.2).<br />

The alpha helix (α-helix) displays a cylindrical structure (Figure 5.2a) where the polypeptidic<br />

backbone possesses a helical conformation strongly folded into the internal cylindrical core.<br />

The side lateral chains are extended out <strong>of</strong> the helical distribution. The α-helix is stabilised by<br />

the hydrogen bonding along the principal backbone chain. Within this, the separation between<br />

the amino acids is close to 1.5 Å, <strong>and</strong> each complete helical turn has 3.6 amino acids<br />

(4)(91)(92). Alanine <strong>and</strong> leucine are strong helix favouring residues, while proline is rarely<br />

found in helices, because in its backbone nitrogen is not to available for the hydrogen bonding<br />

required for helix formation. The aromatic side chain <strong>of</strong> phenylalanine can sometimes<br />

participate in weakly polar interactions (4).<br />

The beta sheet fold (β-sheet) differs markedly from the α-helix. A polypeptide chain, called a β-<br />

str<strong>and</strong>, in a β-sheet is almost fully extended rather than being tightly coiled as in the α-helix.<br />

(Figure 5.2b). The distance between adjacent amino acids along the β-str<strong>and</strong>s is approximately<br />

3.5 Å, in contrast with the distance <strong>of</strong> 1.5 Å along the α-helix. A β-sheet is formed by linking<br />

two or more β-str<strong>and</strong>s by hydrogen bonds. Adjacent chains in β-sheets can run in opposite<br />

direction (antiparallel β-sheet) or in the same direction (parallel β-sheet). In the antiparallel<br />

arrangement, the –NH <strong>and</strong> the –CO groups <strong>of</strong> each amino acid are, respectively, hydrogen<br />

bonded to the –CO <strong>and</strong> the –NH groups <strong>of</strong> a partner on the adjacent chain. In the parallel<br />

arrangement, the –NH group is hydrogen bonded to the –CO group <strong>of</strong> one amino acid on the<br />

129


adjacent str<strong>and</strong>, whereas the –CO group is hydrogen bonded to the –NH group on the amino<br />

acid two residues farther along the chain (4)(92).<br />

Figure 5.2 Secondary structures <strong>of</strong> the proteins a) -helix, b) β-sheet <strong>and</strong> c) turns.<br />

In proteins, the directional changes <strong>of</strong> the polypeptide chains are led by simple structural<br />

elements like inverse turn (Figure 5.2c), called β-turns or hairpin turns <strong>and</strong> loops. This<br />

structure stabilized abrupt direction changes <strong>of</strong> the polypeptide chains. Contrary to α-helix <strong>and</strong><br />

β-sheet, turns <strong>and</strong> loops have no periodic structures (turns have a compact folded <strong>and</strong> usually<br />

rigid structure stabilized by hydrogen bonding, while loops have an extended <strong>and</strong> disorder<br />

structure without hydrogen bonds). These are the connecting elements that link successive<br />

runs <strong>of</strong> the α-helix or β-sheet conformations. Particularly, β-turns connect the ends <strong>of</strong> two<br />

130


adjacent segments <strong>of</strong> an antiparallel β-sheet. The structure is a 180 0 turn involving four amino<br />

acid residues, with a carbonyl oxygen <strong>of</strong> the first residue forming a hydrogen bond with the<br />

amino-group hydrogen <strong>of</strong> the fourth. The peptide groups <strong>of</strong> the central two residues do not<br />

participate in any inter-residue hydrogen bonding (Gly <strong>and</strong> Pro residues <strong>of</strong>ten occur in β-turn,<br />

the former because it is small <strong>and</strong> flexible, the later because peptide bounds involving the<br />

amino nitrogen <strong>of</strong> proline readily assume cis configuration). Generally, β-turns are <strong>of</strong>ten<br />

localized near the surface <strong>of</strong> a protein, where the peptide groups <strong>of</strong> the central two amino acid<br />

residues in the turn can hydrogen-bonded with water. Hence, the secondary structures <strong>of</strong><br />

proteins make reference to highly organized regular geometries. Turns <strong>and</strong> loops invariably lie<br />

on the surfaces <strong>of</strong> proteins <strong>and</strong> thus participate in interactions between proteins <strong>and</strong> other<br />

molecules (91)(92).<br />

5.1.3.- Tertiary Structure <strong>of</strong> the proteins<br />

The overall three-dimensional arrangement <strong>of</strong> atoms in a protein is referred to as the protein<br />

tertiary structure. Whereas the secondary structure refers to the spatial arrangement <strong>of</strong> amino<br />

acid residues that are adjacent in a segment <strong>of</strong> a polypeptide, tertiary structure includes<br />

longer-range aspects <strong>of</strong> the amino acid sequence. Amino acids that are far apart in the<br />

polypeptide sequence <strong>and</strong> are in different types <strong>of</strong> secondary structure may interact within the<br />

completely folded structure <strong>of</strong> a protein. In other words, tertiary structure emerges from the<br />

distribution <strong>of</strong> side chains <strong>of</strong> amino acid residues. Instead, the buried parts <strong>of</strong> proteins consist<br />

almost entirely <strong>of</strong> non-polar residues. On the other h<strong>and</strong>, charged residues usually are absent<br />

from the inside <strong>of</strong> proteins. However, both polar <strong>and</strong> non-polar residues can be found on the<br />

protein surface. In aqueous environment, protein folding is driven by the strong tendency <strong>of</strong><br />

hydrophobic residues to be excluded from water (the system is thermodynamically stable<br />

when hydrophobic groups are clustered rather than extended into the aqueous surroundings).<br />

The polypeptide chain therefore folds so that its hydrophobic side chains are buried <strong>and</strong> its<br />

polar, charged chains are on the surface. Recall that many α-helix <strong>and</strong> β-str<strong>and</strong>s are<br />

amphipatic since both secondary structures have a hydrophobic face, which points into the<br />

protein interior, <strong>and</strong> a more polar face, which points into solution. The buried domains in a<br />

hydrophobic environment are stabilized by hydrogen bonding, pairing all the amino <strong>and</strong><br />

carboxylic groups. This pairing is neatly accomplished in an α-helix <strong>and</strong> β-sheet structures. Van<br />

der Waals interactions between tightly packed hydrocarbon side chains also contribute to the<br />

stability <strong>of</strong> the tertiary protein structure (4).<br />

131


5.1.4.- Quaternary structure <strong>of</strong> the proteins<br />

The quaternary structure refers to the association <strong>of</strong> several polypeptide chains (with tertiary<br />

structure) linked by weak forces. Amongst the forces stabilising the quaternary structure we<br />

find hydrogen bonding, hydrophobic interactions <strong>and</strong> cross-linking interactions like disulfide<br />

bridges <strong>and</strong> metal-ion ligation. At this structural level, the protein assemblies composed <strong>of</strong><br />

more than one polypeptide chain are called oligomers, <strong>and</strong> each individual chain is called<br />

subunit.<br />

5.2.- Protein aggregation<br />

The timescale required to protein folding in biological systems is quickly enough to avoid a<br />

r<strong>and</strong>om mechanism for protein folding (93). As a general principle, the kinetic pathway <strong>of</strong><br />

protein folding can be described as an ordered fold reaction. In this way, the folding <strong>of</strong> only<br />

very simple proteins has been determined at atomic level resolution to date (94), using<br />

biophysical <strong>and</strong> theoretical experiments as a sequential <strong>and</strong> highly cooperative reaction (95).<br />

The concept <strong>of</strong> energy l<strong>and</strong>scape is used to visualize the conformational space available to<br />

each individual polypeptide under a given condition. This theoretical formalism describes the<br />

progressive folding <strong>of</strong> polypeptide chains along their energetic structural evolution (the energy<br />

l<strong>and</strong>scape) from an unfolded conformation to a compact native structure (Figure 5.3) (96). For<br />

small proteins, this l<strong>and</strong>scape appears to be funnel-like <strong>and</strong> represent the evolutionary<br />

selection <strong>of</strong> a polypeptide sequence able to fold rapidly <strong>and</strong> reliably towards a unique native<br />

state. On the other h<strong>and</strong>, larger polypeptide sequences have rougher energy l<strong>and</strong>scapes, in<br />

which there are present different populations <strong>of</strong> partially folded species that may be on- or<br />

<strong>of</strong>f-pathway to the native fold. Characterizing the multitude <strong>of</strong> populated conformational<br />

states on this energy l<strong>and</strong>scape is not only crucial for developing an underst<strong>and</strong>ing <strong>of</strong> the<br />

determinants <strong>of</strong> protein folding <strong>and</strong> function, but also contains crucial clues about side-<br />

reactions such as protein aggregation. Such characterization involves the knowledge <strong>of</strong> all<br />

structural, kinetic <strong>and</strong> thermodynamic properties <strong>of</strong> all conformations accessible to a given<br />

polypeptide sequence. However, it is necessary to develop analytical-high resolution tools<br />

capable to resolve the structural fast interconversion <strong>of</strong> the species involved, as well as the<br />

theoretical bases that permit to underst<strong>and</strong> <strong>and</strong> interpret the obtained data (97).<br />

132


5.2.1.- Unfolded state<br />

Over the last decade, an enormous progress has been made in the description <strong>of</strong> the<br />

conformation <strong>of</strong> the unfolded states <strong>of</strong> polypeptide chains. These important species not only<br />

define the starting point for protein folding, but also have important roles in a variety <strong>of</strong><br />

biological process. When described the unfolded state <strong>of</strong> a polypeptide sequence, the first idea<br />

in mind is a r<strong>and</strong>om coil structure, which lacks <strong>of</strong> specific inter-residue interactions between<br />

the lateral chains <strong>of</strong> the amino acid sequence. However, this viewpoint has moved over recent<br />

years as a result <strong>of</strong> the increasing number <strong>of</strong> detailed studies about the properties <strong>and</strong><br />

structure <strong>of</strong> denatured states under different solution conditions (98). Current views consider<br />

that even in the presence <strong>of</strong> high denaturant concentrations, significant amounts <strong>of</strong> aliphatic<br />

<strong>and</strong> aromatic side chains present in the native state may persist, even when the secondary<br />

structure <strong>of</strong> the native state has been lost. On the other h<strong>and</strong>, under lesser harsh solution<br />

conditions, also some proteins can be unfolded in the absence <strong>of</strong> denaturant agents (for<br />

example, by mutating the sequence or changing the pH <strong>of</strong> the solution), the unfolded state has<br />

been shown to contain significant native-like interactions with substantial native <strong>and</strong> non-<br />

native side chain contacts which allow the existence <strong>of</strong> a residual structure. This residual<br />

structure in the unfolded state has not only been suggested to reduce the conformational<br />

search during protein folding, but also for some proteins may play a role in the onset <strong>of</strong><br />

protein aggregation <strong>of</strong> unfolded polypeptide sequences (97)(99).<br />

5.2.2.- Intermediate <strong>and</strong> transition state ensembles<br />

The function <strong>of</strong> the intermediate states (populated partially-folded states) during protein<br />

folding has been a long-st<strong>and</strong>ing question. Supported by the observation that many small<br />

proteins fold with apparent two-state kinetics, folding intermediates were initially thought to<br />

be aberrant misfolds on the folding energy l<strong>and</strong>scape, representing <strong>of</strong>f-pathway species or<br />

kinetics traps. Nowadays, thanks to more powerful experimental methods (Foster resonance<br />

emission transfer fluorescence (FRET), fluorescent correlation spectroscopy or NMR<br />

spectroscopy, amongst others) protein folding studies have revealed that partially folded<br />

states are widely populated during the protein folding process, even in simplest proteins (97).<br />

In this regard, a model that involves transient intermediate states in equilibrium into the<br />

pathway <strong>of</strong> a given polypeptide chain has been suggested. Under this context, a polypeptide<br />

chain with high molecular weight may present many intermediate states, describing a<br />

roughness energy l<strong>and</strong>scape for its particular folding pathway. For example, larger polypeptide<br />

chains have a higher tendency to collapse under specific solution conditions, leading to the<br />

formation <strong>of</strong> compact states, which can contain substantial native-like structure. Structural<br />

133


eorganization as a consequence <strong>of</strong> inter-residue contact (including both native <strong>and</strong> non<br />

native-like interactions) <strong>of</strong> these compact states may involve high-energy barriers, leading to<br />

the transient population <strong>of</strong> partially intermediate states. Such species can be either productive<br />

for on-pathway folding until reaching the native state; or they can be trapped in such a manner<br />

that the native structure cannot be reached without substantial organizational events (<strong>of</strong>f-<br />

pathway).<br />

5.2.3.- Native state<br />

Currently, it is known that the conformational states <strong>of</strong> proteins are structural dynamic<br />

processes, i.e. even at native conditions, structured proteins have access to a manifold <strong>of</strong> near-<br />

native conformations, in which protein native structures show fluctuations around the minimal<br />

energy conformation (100). Such structural movements are essential in order to encompass<br />

their function, such as enzyme catalysis or lig<strong>and</strong> binding, as well as site-to site communication<br />

within the globular protein fold or between proteins subunits.<br />

Figure 5.3 Schematic illustration <strong>of</strong> the energy l<strong>and</strong>scape for protein folding <strong>and</strong> aggregation.<br />

a) The surface illustrates the roughness <strong>of</strong> the protein l<strong>and</strong>scape, showing the multitude <strong>of</strong><br />

conformational states available to a polypeptide chain. The conformational search <strong>of</strong> a single<br />

polypeptide chain to a functional monomer can be described in a simple folding funnel (light<br />

grey), intermolecular association increased the roughness <strong>of</strong> the energy l<strong>and</strong>scape (dark grey),<br />

b) Proposed pathways related to the conformational states shown in a) populated on the<br />

combined folding <strong>and</strong> aggregation energy l<strong>and</strong>scape (97).<br />

134


5.2.4.- Energy l<strong>and</strong>scape: protein aggregation<br />

The schematic diagram in Figure 5.3 attempts to depict the complexity <strong>of</strong> the protein folding<br />

<strong>and</strong> aggregation energy l<strong>and</strong>scapes as previously mentioned; at the same time, it makes an<br />

effort to illustrate the wide variety <strong>of</strong> different conformational states with minimum energies<br />

<strong>and</strong> the large diversity <strong>of</strong> folding pathways available for each polypeptide chain as it<br />

circumnavigates the l<strong>and</strong>scape. Regarding the in-depth knowledge <strong>of</strong> the folding l<strong>and</strong>scape <strong>of</strong><br />

simple, single domain monomeric proteins, <strong>and</strong> the conformational states accessible to<br />

polypeptide oligomers, relatively little is understood. Energy minima on the aggregation side <strong>of</strong><br />

the energy l<strong>and</strong>scape might be poorly defined, as a result <strong>of</strong> broad ensembles <strong>of</strong> oligomeric<br />

states <strong>of</strong> similar energy that are rapidly interconverting, but could also be highly defined, as<br />

might be expected for higher order species such prot<strong>of</strong>ibrils or fibrils. For example, the energy<br />

minimum <strong>of</strong> mature amyloid fibrils might be deeper <strong>and</strong> sharper that those <strong>of</strong> native<br />

monomeric proteins, as suggested by the rigidity <strong>of</strong> the amyloid fold, as well as by the<br />

nucleation-dependent polymerisation mechanism (101). Here, however, even under the same<br />

solution conditions a multitude <strong>of</strong> fibril morphologies can be formed simultaneously,<br />

highlighting the complexity <strong>and</strong> multiplicity <strong>of</strong> the aggregation pathway (102)(103).<br />

5.3.- Amyloid fibrils<br />

Into the context <strong>of</strong> the energy l<strong>and</strong>scape, the observation that virtually any protein sequence<br />

can form amyloid fibrils given the appropriate solution conditions led to the suggestion that<br />

the amyloid fold is the universal global free energy minimum <strong>of</strong> all polypeptide chains that<br />

may assemble by generic mechanisms governed by the physicochemical properties <strong>of</strong> the<br />

polypeptide chain.<br />

5.3.1. Amyloid fibrils hallmarks<br />

Amyloid comes from the Greek word amylon which means starch. Amyloid was coined initially<br />

in a botanical context by Scheiden (104). Virchow <strong>and</strong> others researches used the term<br />

amyloid into medicine to describe human pathogenic deposits that stain with sulphuric acid<br />

<strong>and</strong> iodine solution, similarly to starch staining. Although these deposits do not primarily<br />

consist <strong>of</strong> polysaccharides, they were found to be a proteinaceus, nowadays the definitions <strong>of</strong><br />

amyloid fibrils depend on the context <strong>of</strong> their use. In pathological diagnosis, amyloid had been<br />

defined as extracellular deposits <strong>of</strong> proteins fibrils with a characteristic appearance in electron<br />

microscope, a typical X-ray diffraction pattern, <strong>and</strong> a strong affinity for Congo red dye (105).<br />

135


In vivo, the self-assembly <strong>of</strong> proteins into highly ordered, β-sheet enriched, fibrillar, insoluble,<br />

supramolecular structures known as amyloid fibrils (or amyloid-like fibrils, term referring to<br />

protein fibrils that not associated with any disease) is linked to the onset <strong>of</strong> a growing number<br />

<strong>of</strong> human disorders, including Alzheirmer’s, Parkinson’s, Huntington´s diseases, spongiform<br />

encephalopathies, <strong>and</strong> type II diabetes mellitus. On the other h<strong>and</strong>, the controlled self-<br />

assembly <strong>of</strong> proteins <strong>and</strong> peptides into amyloid-like structures may constitute an attractive<br />

alternative to develop nanomaterials (106)(107).<br />

Amyloid fibrils can be obtained in vitro through different ways by adjusting different external<br />

parameters as, for example, pH, ionic strength, <strong>and</strong> temperature, <strong>and</strong> cosolute or cosolvent<br />

additions. This means that the formation <strong>of</strong> amyloid-like conformations is not restricted to<br />

peptide chains <strong>of</strong> specific length or having a particular primary or secondary structure; this<br />

suggests that the conformational shift into β-sheet rich is energetically preferred. The<br />

characterization <strong>of</strong> the fibrillation process <strong>of</strong> proteins in vitro has largely been focused on<br />

biophysical characterization in order to determine the structure <strong>of</strong> the fibril, <strong>and</strong> on<br />

biochemical studies to get into the mechanism <strong>and</strong> kinetics <strong>of</strong> the process. Also, through<br />

techniques such X-ray diffraction, electron microscopy, <strong>and</strong> solid state NMR spectroscopy we<br />

now have a good underst<strong>and</strong>ing <strong>of</strong> the core architecture <strong>of</strong> individual fibrils too.<br />

In general, amyloid fibrils possess distinct physical properties as: i) they are characterized by a<br />

cross-β sheet structure, where individual β-str<strong>and</strong>s are perpendicular <strong>and</strong> each β-sheet is<br />

parallel to the fibril axis (Figure 5.4). The X-ray diffraction pattern shows two characteristic<br />

reflections at 4.7 Å <strong>and</strong> 10-11 Å, corresponding to the inter-str<strong>and</strong> <strong>and</strong> stacking distances<br />

between individual β-sheets, respectively; ii) amyloids bind to histological dyes such as Congo<br />

red (CR) <strong>and</strong> Thi<strong>of</strong>lavin T (ThT). After binding to CR, an apple-green birefringence under cross-<br />

polarized light is observed. CR <strong>and</strong> ThT dyes, however, do not bind to monomeric proteins or<br />

peptides. These dyes fluoresce when they bind to β-sheet rich fibrils <strong>and</strong> are, therefore, useful<br />

for spectroscopic monitoring <strong>of</strong> fibril´s growth <strong>and</strong> kinetics; iii) under electron microscope<br />

(EM), amyloid fibrils appear to be few micrometers long, non-branched filaments, with 7-12<br />

nm diameter. In a majority <strong>of</strong> cases, amyloids are composed <strong>of</strong> 2-4 prot<strong>of</strong>ilaments that are<br />

either helically twisted or laterally associated with each other forming higher order fibrils; iv)<br />

amyloids are resistant to heat, wide ranges <strong>of</strong> pH <strong>and</strong> proteases (107)-(113).<br />

Over the past several years, soluble proteins <strong>and</strong> protein fragments self-assembled into<br />

amyloid aggregates have attracted the interest <strong>of</strong> researchers from diverse fields. From a<br />

136


molecular medicine viewpoint, protein deposition is an undesired process associated with<br />

several human diseases. Hence, different questions must be arisen such as why self-<br />

association <strong>of</strong> a misfolded protein results in organ dysfunction <strong>and</strong> neurodegeneration?, what<br />

are the toxic species?, <strong>and</strong> whether aggregates <strong>of</strong> all proteins are toxic or not?. At the<br />

molecular level, fundamental research has been directed towards obtaining a rational<br />

underst<strong>and</strong>ing <strong>of</strong> the physico-chemical principles underlying protein misfolding <strong>and</strong> self-<br />

assembly. However, given the complexity <strong>of</strong> the molecular mechanisms driving amyloid<br />

aggregation, a frequently used strategy in the field consists <strong>of</strong> design polypeptide model<br />

systems to allow biophysical characterization <strong>and</strong> the study <strong>of</strong> their aggregation process (114)-<br />

(118). This strategy permits the analysis <strong>of</strong> the relationship between the amino acid sequence<br />

<strong>of</strong> the polypeptide chain <strong>and</strong> its structural stability during the conformational transition.<br />

Figure 5.4 Schematic representation <strong>of</strong> an amyloid fibril structure: a) β-str<strong>and</strong>s perpendicular<br />

to the fibre axis; the β-sheet structure arises from the hydrogen bonding between lateral side<br />

chains <strong>of</strong> amino acids; b) cross-β diffraction pattern observed in amyloid fibrils. Arrows<br />

indicate the positions <strong>of</strong> the cross-β reflection on the meridian at 4,7 Å (white) <strong>and</strong> on the<br />

equator at 10 Å (black) (112).<br />

a) b)<br />

Several studies have shown that the destabilization <strong>of</strong> the polypeptide sequence per se is not<br />

enough to bring the self-assembly process up to form amyloid fibrils (119)(120). Additionally,<br />

the sudden conformational transition undergone by proteins in their native state to form<br />

amyloid fibrils rich in β-sheets is favoured only if the amino acid sequence increase its ability to<br />

form β-sheet structures. In most <strong>of</strong> these systems, the structural transitions from α-helix to β-<br />

structures are originated by applying external heat energy. This phenomenon involves a<br />

number <strong>of</strong> relatively weak non-covalent molecular forces: hydrophobic, electrostatic, <strong>and</strong><br />

137


hydrogen bonds mainly. Molecular simulations (109) demonstrated that hydrophobic<br />

interactions play an important role in promoting protein self-assembly. Meanwhile, for small<br />

peptides the β-sheet aggregates are stabilized by backbone hydrogen bonds, as well as by<br />

specific side-chain interactions such as hydrophobic stacking <strong>of</strong> polar side chains <strong>and</strong><br />

electrostatic interaction.<br />

5.3.2.- Kinetic fibril formation in vitro<br />

The generic nature <strong>of</strong> the aggregation process has enabled extensive studies <strong>of</strong> the transition<br />

between soluble precursor states <strong>and</strong> insoluble amyloid fibril in vitro (121). The aggregation<br />

into amyloid structure <strong>of</strong> globular proteins has been shown to require a conformational<br />

transition process before reaching the amyloid fibre structure. To carry out this process, the<br />

native conformation <strong>of</strong> globular proteins usually must be destabilised by the addition <strong>of</strong><br />

denaturants agents, low pH, high temperature, truncation/mutation <strong>and</strong> so (122)-(124).<br />

Specific protein destabilisation results in an increased population <strong>of</strong> partially folded<br />

conformations exposing aggregation-prone regions that are usually protected in the native<br />

state <strong>and</strong>, thus, enhancing the probability <strong>of</strong> non-native intermolecular interactions.<br />

At present, the proposed models to describe the formation process <strong>of</strong> amyloid fibrils include<br />

monitoring the fibrillogenesis by high-resolution microscopic <strong>and</strong> spectroscopic tools, <strong>and</strong> by<br />

the binding <strong>of</strong> specific dyes to amyloid fibrils. The most common model for explaining the<br />

amyloid formation is the nucleation-dependent polymerization model. In such a model, the<br />

kinetic pathway from the native structure to amyloid fibrils occurs by a nucleation-aggregation<br />

process. The nucleation step (also known as lag phase) is the factor that determines the rate <strong>of</strong><br />

amyloid formation by the slow generation <strong>of</strong> nuclei under successive thermodynamically<br />

unfavourable steps; these nuclei aggregate when a critical protein concentration value has<br />

been exceeded by a small amount; hence, the process is concentration-dependent. The lag<br />

phase precedes a period <strong>of</strong> exponential growth that corresponds to fibril elongation,<br />

characterised by a sudden fibre growth as a consequence <strong>of</strong> the addition <strong>of</strong> protein monomers<br />

or the association <strong>of</strong> other competent species; this step is called elongation phase. In contrast,<br />

another proposed model <strong>of</strong> fibrillogenesis is the “downhill polymerization” process, which<br />

consists <strong>of</strong> a continuum aggregate state without requiring a nucleation step. A relatively easy<br />

way to observe the fibrillation kinetics is by using an amyloid specific binding dye such<br />

Thi<strong>of</strong>lavin T (ThT). ThT increases its fluorescence intensity upon binding to amyloid fibrils when<br />

using an excitation wavelength <strong>of</strong> 440 nm, <strong>and</strong> collecting the emission intensity at 480 nm.<br />

Figure 5.5 shown the kinetic formation <strong>of</strong> amyloid fibril followed by ThT binding. Figure 5.5 a)<br />

138


shows the typical nucleation-dependent polymerization characterised by a sudden increase <strong>of</strong><br />

<strong>of</strong> ThT fluorescence as a function time; once the lag phase is overcome (the lag phase is<br />

observed as the region where no appreciable changes in ThT fluorescence intensity take<br />

place). In contrast, with the non nucleation-dependent polymerization (downhill<br />

polymerization, Figure 5.5b) shows a marked <strong>and</strong> continuous ThT fluorescence intensity<br />

increase (125)-(129).<br />

In vitro, the aggregation process <strong>of</strong> polypeptide chains into fibrillar structures can be also<br />

monitored by means <strong>of</strong> other analytical tools such as circular dichroism (CD), infrared<br />

spectroscopy (IR), nuclear magnetic resonance spectroscopy (NMR), light scattering,<br />

fluorescence spectroscopy (ThT), or imaging techniques such as TEM, SEM or AFM.<br />

Figure 5.5 ThT fluorescence intensity as a function <strong>of</strong> time following a: a) nucleation-<br />

dependent polymerization, <strong>and</strong> b) a non-nucleation-dependent polymerization process.<br />

5.4.- Human serum albumin (HSA)<br />

HSA is a globular protein constituted by 585 amino acids in a single polypeptide chain with a<br />

molecular weight <strong>of</strong> 67,000 Da. Human serum albumin is the most abundant protein in blood<br />

plasma, it plays a special role in transporting endogenous <strong>and</strong> exogenous metabolites through<br />

the vascular system <strong>and</strong> also maintains the pH <strong>and</strong> osmotic pressure <strong>of</strong> plasma (130). The<br />

protein structure contains 17 disulfide bridges giving rigidity <strong>and</strong> stability to the protein<br />

structure. HSA has one free –SH group, which leads to protein dimerization <strong>and</strong> high order<br />

association. Accordingly to X-ray crystallography data, the secondary structure <strong>of</strong> HSA<br />

molecule is composed <strong>of</strong> 67% -helix, no -sheet, 10% turn, <strong>and</strong> 23% extended chain<br />

(131)(132). On the other h<strong>and</strong>, in agreement with CD spectroscopy data, the secondary<br />

139


structure <strong>of</strong> HSA molecule at physiological solution conditions is composed <strong>of</strong> 67% -helix, 5%<br />

-sheet, 12% turn, <strong>and</strong> 17% extended chain (133). The three-dimensional configuration <strong>of</strong><br />

serum albumin is divided into three major domains (labelled as I, I <strong>and</strong> III), <strong>and</strong> each domain<br />

are composed <strong>of</strong> two subdomains that possess common structural motifs, which are<br />

predominantly helical <strong>and</strong> cross-linked by the disulfide bridges (134)(135).<br />

The HSA structure can be influenced by several factors as pH, temperature, cosolvents... It is<br />

known that HSA molecule shows a number <strong>of</strong> well-defined pH-dependent structural<br />

transitions, (Figure 5.6) (136): E form (below pH 2.7), F form (pH 2.7-4.3), N form (pH 4.3-8), B<br />

form (pH 8-10) <strong>and</strong> A form (over pH 10) (134)-(137). On the other h<strong>and</strong>, during its thermal<br />

unfolding process, three states can be distinguished: the native form, the reversible unfolded<br />

form (at 65 0 C) <strong>and</strong> the irreversible unfolded form (at 74 0 C) (120). Both unfolding processes<br />

give rise to aggregation-prone conformational changes <strong>of</strong> both secondary <strong>and</strong> tertiary<br />

structures, allowing the formation <strong>of</strong> amorphous aggregates, intermediate states, oligomers<br />

<strong>and</strong> amyloid fibrils.<br />

Due to the physiological importance <strong>of</strong> human serum albumin as a carrier protein <strong>and</strong> blood<br />

pressure regulator <strong>and</strong> its propensity to easily aggregate in vitro, HSA has become a good<br />

model for protein aggregation studies. Moreover, as the phenomenon <strong>of</strong> protein aggregation<br />

appears to reflect certain generic “polymeric” features <strong>of</strong> proteins (138), the study <strong>of</strong> protein<br />

aggregation mechanisms in model systems is extremely useful to gain a better underst<strong>and</strong>ing<br />

about the molecular mechanisms <strong>of</strong> disease-associated amyloidogenesis.<br />

In this way, in this work we show the propensity <strong>of</strong> HSA to form amyloid fibrils <strong>and</strong> other types<br />

<strong>of</strong> aggregates under different thermal <strong>and</strong> solvent conditions (changing the solution pH,<br />

solution ionic strength <strong>and</strong> solvent composition), allowing the modulation <strong>of</strong> electrostatic,<br />

hydrophobic <strong>and</strong> hydration interactions between protein molecules <strong>and</strong> aggregates (139)-<br />

(141). The kinetic aggregation, protein conformational changes upon self-assembly, <strong>and</strong><br />

structure <strong>of</strong> the different intermediates on the fibrillation pathway were determined by means<br />

<strong>of</strong> ThT fluorescence <strong>and</strong> CR absorbance; far <strong>and</strong> near-UV-Vis CD; tryptophan fluorescence;<br />

Fourier transform infrared spectroscopy; X-ray diffraction; <strong>and</strong> TEM, SEM, AFM <strong>and</strong> optical<br />

microscopy. We have observed that the fibril formation is largely affected by electrostatic<br />

shielding: at physiological pH, fibrillation is progressively more efficient <strong>and</strong> faster in the<br />

presence <strong>of</strong> up to 50 mM NaCl; meanwhile, at larger salt concentrations, excessive charge<br />

shielding <strong>and</strong> further enhancement <strong>of</strong> the solution hydrophobicity might involve a change in<br />

140


the energy l<strong>and</strong>scape <strong>of</strong> the aggregation process, which makes the fibrillation process difficult.<br />

In contrast, under acidic conditions a continuous progressive enhancement <strong>of</strong> HSA fibrillation<br />

is observed as the electrolyte concentration in solution increases. Both the distinct protein<br />

ionization state <strong>and</strong> the initial structural states <strong>of</strong> the protein before incubation may be the<br />

origin <strong>of</strong> this behavior. According to ThT fluorescence data, the fibrillation kinetics <strong>of</strong> HSA does<br />

not show a lag phase except at pH 3.0 in the absence <strong>of</strong> added salt. The HSA fibrils obtained at<br />

the end <strong>of</strong> the fibrillation process show structural features typical <strong>of</strong> classical amyloid fibers, as<br />

denoted by XRD, CD <strong>and</strong> TEM techniques. Finally, we describe the physical <strong>and</strong> structural<br />

properties <strong>of</strong> additional supraself-assembled structures <strong>of</strong> HSA under solution conditions in<br />

which amyloid fibrils are formed. We have detected the formation <strong>of</strong> ordered aggregates <strong>of</strong><br />

amyloid fibrils, i.e. spherulites, by means <strong>of</strong> polarized optical microscopy, laser confocal<br />

microscopy <strong>and</strong> TEM. These structures possess a radial arrangement <strong>of</strong> the fibrils around a<br />

disorganized protein core, <strong>and</strong> have sizes <strong>of</strong> several micrometers.<br />

Figure 5.6 pH conditions <strong>and</strong> helical contents <strong>of</strong> the five organized forms <strong>of</strong> albumin including<br />

crystal structure <strong>of</strong> the N form, <strong>and</strong> the proposed configuration <strong>of</strong> the F <strong>and</strong> E forms (136).<br />

5.5.- Amyloid fibrils as biomaterials<br />

The bottom-up strategy takes advantage <strong>of</strong> the functionality <strong>of</strong> molecular materials <strong>and</strong><br />

molecular recognition properties for self-assembly (142). The construction <strong>of</strong> functional<br />

nanomaterials <strong>and</strong> components through the bottom-up strategy has potential applications in<br />

different areas as electronics, tissue engineering, drug delivery, sensing… (143)(144). In this<br />

regard, self-assembly properties <strong>of</strong> biomacromolecules, as phospholipids, DNA, peptides, <strong>and</strong><br />

proteins, can be an alternative for the obtention <strong>of</strong> new materials.<br />

141


The physical <strong>and</strong> chemical properties <strong>of</strong> proteins are an attractive alternative for the<br />

construction <strong>of</strong> nanoestructured materials, providing an extraordinary array <strong>of</strong> functionalities<br />

that could be exploited in fields such as biomedicine, biotechnology, materials science, <strong>and</strong><br />

nanotechnology (145). Protein amyloids are excellent c<strong>and</strong>idates for the fabrication <strong>of</strong><br />

molecular nanobiomaterials, such wires, layers, gels, scaffolds, templates, <strong>and</strong> liquid crystals,<br />

using the bottom-up strategy as a result <strong>of</strong> their outst<strong>and</strong>ing physico-chemical properties<br />

(great stiffness, stable against heat <strong>and</strong> denaturants, resistant to proteases, amongst others)<br />

structural compatibility, nanoscale dimensions <strong>and</strong> efficient assembly into well-defined<br />

ultrastructures (146)(147).<br />

On the other h<strong>and</strong>, metal nanoparticles (NP) have received considerable attention during last<br />

decade because <strong>of</strong> their particular optical, electronic, magnetic <strong>and</strong> catalytic properties <strong>and</strong><br />

their important applications in many fields such as nanosensors, catalysis, biomedicine,<br />

biological, labelling, <strong>and</strong> surface-enhanced Raman scattering (SERS). To optimize <strong>and</strong> extend<br />

the applications <strong>of</strong> metal NPs, methods must be developed to control the assembly <strong>and</strong><br />

organization <strong>of</strong> these nanomaterials. NPs in ordered arrays provide optical <strong>and</strong> electronic<br />

properties that are distinct compared to individual particles or disorganized macroscale<br />

agglomerations. In this way, the supramolecular structure <strong>of</strong> amyloid fibrils can be used to<br />

organize nanoparticles into predefined, topologically intricate nanostructures, or synthesize<br />

miscellaneous materials in order to control the properties <strong>of</strong> nanoparticle assemblies for<br />

potential applications in electronic, optical, <strong>and</strong> chemical devices. In particular, functionalizing<br />

one-dimensional (1D) supporting biomaterials with metal NPs that combine the properties <strong>of</strong><br />

two functional materials, such as the high conductivity, surface area or the precise chemical<br />

functionality <strong>of</strong> the biotemplate, <strong>and</strong> the unique plasmonic or catalytic properties <strong>of</strong> the metal<br />

NPs widen their range <strong>of</strong> applications <strong>and</strong>, therefore, their important role in nanoscience <strong>and</strong><br />

nanotechnology (148)-(151).<br />

5.5.1 Hen egg-white lysozyme<br />

Hen egg-white lysozyme (HEWL) is a relatively small globular, monomeric protein. HEWL<br />

consist <strong>of</strong> a polypeptide chain with 129 amino acid residues with four disulfide bonds, <strong>and</strong> it<br />

folds into two structural domains, the α <strong>and</strong> β ones. The α domain contains a core <strong>of</strong><br />

hydrophobic side chains that are packed closely together (152); this domain consists <strong>of</strong> three<br />

long helices A (5-15), B (25-35) <strong>and</strong> C (89-99) (153). In contrast, the β domain <strong>of</strong> the protein<br />

does not display a similar hydrophobic core; instead, hydrogen bonds <strong>and</strong> a number <strong>of</strong> small<br />

hydrophobic clusters appear to be responsible for defining its tertiary fold, <strong>and</strong> there is also an<br />

142


exposed loop region. The secondary structure <strong>of</strong> lysozyme dissolved in water is composed <strong>of</strong><br />

30% -helix, 18% -sheet, 22% turn, <strong>and</strong> 30% unordered structure (153)-(155). It has been also<br />

found that the HEWL easily forms amyloid fibrils under several denaturant solution conditions,<br />

such acidic conditions <strong>and</strong> elevated temperature (152)(155). In this regard, it has been<br />

proposed that fragment 57-107 is a highly amyloidogenic region, <strong>and</strong> probably the conversion<br />

to the β-sheet structure <strong>of</strong> the C subdomain is involved in the amyloid fibril formation rather<br />

than the other subdomains A <strong>and</strong> B (152)(153).<br />

It is known that the proteins have multiple potential ion-binding sites within its aminoacid<br />

sequence; therefore, protein fibrils can be used in a metallization reaction under relative harsh<br />

conditions. In this regard, we take advantage about the spontaneous self-assembly <strong>of</strong> both<br />

HSA <strong>and</strong> HEWL into amyloid fibrils under fibrillation solution conditions, which enables the use<br />

<strong>of</strong> these fibrils as a template for the construction <strong>of</strong> 1-D hybrid material, in particular, gold <strong>and</strong><br />

Fe3O4 nanowires. The inherent structural differences between HSA <strong>and</strong> HEWL fibrils, such as<br />

strength, length <strong>and</strong> stability make lysozyme (Lys) fibers (with widths <strong>of</strong> 4-10 nm, lengths 0.1-2<br />

μm) most suitable for their use in the obtention <strong>of</strong> 1-D nanowires.<br />

In this work, gold nanowires were obtained by seeded-growth process. The degree <strong>of</strong> metal<br />

coverage <strong>of</strong> the biotemplate was controlled by sequential addition <strong>of</strong> salt gold growth<br />

solutions. The hybrid metallic fibrils have been proved to be useful as catalytic substrates,<br />

providing a superior catalytical activity when they are incorporated in the reduction reaction <strong>of</strong><br />

p-nitrophenol to p-aminophenol in the presence <strong>of</strong> NaBH4. The reaction rates obtained were<br />

much larger than those reported for other gold hybrid materials such as Au nanoparticle<br />

networks, PAMAM-dendrimer-suported spherical nNPs, solid 1-D Au nanobelts, <strong>and</strong> nanorods,<br />

polymer micelles-supported Au NPs or 1-D assemblies <strong>of</strong> Au NPs (156)-(159). On the other<br />

h<strong>and</strong>, to obtain iron oxide wires, we in situ synthesize Fe3O4 NPs by precipitation in an alkaline<br />

medium by using the electrostatic interaction <strong>of</strong> iron ions (Fe +2 <strong>and</strong> Fe +3 ) <strong>and</strong> the negative<br />

charged amino acid residues (mainly aspartic <strong>and</strong> glutamic acid) located on the protein fibril<br />

surfaces. Complete magnetic coating on the protein fibril surface was obtained by sequential<br />

nanoprecipitation in situ <strong>of</strong> Fe +2 <strong>and</strong> Fe +3 ions (1:2 molar ratio), in the presence <strong>of</strong> ammonium<br />

hydroxide (NH4OH). Because <strong>of</strong> their large magnetic properties, the 1-dimensional magnetic<br />

nanowires can serve as efficient magnetic resonance image (MRI) contrast agents, with<br />

transverse relaxativities larger than those obtained previously for other 1-D nanostructures<br />

derived by alternative methods (160)-(162).<br />

143


Contents<br />

6.1<br />

6.2<br />

6.3<br />

6.4<br />

6.5<br />

6.6<br />

6.7<br />

6.8<br />

CHAPTER 6<br />

Papers on amyloid fibrils<br />

Existence <strong>of</strong> different structural intermediates on the<br />

fibrillation pathway <strong>of</strong> human serum albumin<br />

Influence <strong>of</strong> electrostatic interactions on the fibrillation<br />

process <strong>of</strong> human serum albumin<br />

Additional supra-self-assembly <strong>of</strong> human serum albumin<br />

under amyloid-like-forming solution conditions<br />

Hydration effects on the fibrillation process <strong>of</strong> globular<br />

protein: The case <strong>of</strong> human serum albumin<br />

Obtention <strong>of</strong> metallic nanowires by protein biotemplating<br />

<strong>and</strong> their catalytic application<br />

One-dimensional magnetic nanowires obtained by protein<br />

fibril biotemplating<br />

Relevant aspects on the HSA fibril formation<br />

Relevant aspects on the gold <strong>and</strong> magnetic nanowires<br />

147<br />

165<br />

175<br />

185<br />

199<br />

207<br />

215<br />

216<br />

145


Biophysical Journal Volume 96 March 2009 2353–2370 2353<br />

Existence <strong>of</strong> Different Structural Intermediates on the Fibrillation Pathway<br />

<strong>of</strong> Human Serum Albumin<br />

Josué Juárez, Pablo Taboada,* <strong>and</strong> Víctor Mosquera<br />

Grupo de Física de Coloides y Polímeros, Departamento de Física de la Materia Condensada, Facultad de Física, Universidad de Santiago de<br />

Compostela, Santiago de Compostela, Spain<br />

ABSTRACT The fibrillation propensity <strong>of</strong> the multidomain protein human serum albumin (HSA) was analyzed under different<br />

solution conditions. The aggregation kinetics, protein conformational changes upon self-assembly, <strong>and</strong> structure <strong>of</strong> the different<br />

intermediates on the fibrillation pathway were determined by means <strong>of</strong> thi<strong>of</strong>lavin T (ThT) fluorescence <strong>and</strong> Congo Red absorbance;<br />

far- <strong>and</strong> near-ultraviolet circular dichroism; tryptophan fluorescence; Fourier transform infrared spectroscopy; x-ray<br />

diffraction; <strong>and</strong> transmission electron, scanning electron, atomic force, <strong>and</strong> microscopies. HSA fibrillation extends over several<br />

days <strong>of</strong> incubation without the presence <strong>of</strong> a lag phase, except for HSA samples incubated at acidic pH <strong>and</strong> room temperature in<br />

the absence <strong>of</strong> electrolyte. The absence <strong>of</strong> a lag phase occurs if the initial aggregation is a downhill process that does not require<br />

a highly organized <strong>and</strong> unstable nucleus. The fibrillation process is accompanied by a progressive increase in the b-sheet (up to<br />

26%) <strong>and</strong> unordered conformation at the expense <strong>of</strong> a-helical conformation, as revealed by ThT fluorescence <strong>and</strong> circular<br />

dichroism <strong>and</strong> Fourier transform infrared spectroscopies, but changes in the secondary structure contents depend on solution<br />

conditions. These changes also involve the presence <strong>of</strong> different structural intermediates in the aggregation pathway, such as<br />

oligomeric clusters (globules), bead-like structures, <strong>and</strong> ring-shaped aggregates. We suggest that fibril formation may take place<br />

through the role <strong>of</strong> association-competent oligomeric intermediates, resulting in a kinetic pathway via clustering <strong>of</strong> these oligomeric<br />

species to yield prot<strong>of</strong>ibrils <strong>and</strong> then fibrils. The resultant fibrils are elongated but curly, <strong>and</strong> differ in length depending on<br />

solution conditions. Under acidic conditions, circular fibrils are commonly observed if the fibrils are sufficiently flexible <strong>and</strong> long<br />

enough for the ends to find themselves regularly in close proximity to each other. These fibrils can be formed by an antiparallel<br />

arrangement <strong>of</strong> b-str<strong>and</strong>s forming the b-sheet structure <strong>of</strong> the HSA fibrils as the most probable configuration. Very long incubation<br />

times lead to a more complex morphological variability <strong>of</strong> amyloid mature fibrils (i.e., long straight fibrils, flat-ribbon structures,<br />

laterally connected fibers, etc.). We also observed that mature straight fibrils can also grow by protein oligomers tending to align<br />

within the immediate vicinity <strong>of</strong> the fibers. This filament þ monomers/oligomers scenario is an alternative pathway to the otherwise<br />

dominant filament þ filament manner <strong>of</strong> the protein fibril’s lateral growth. Conformational preferences for a certain pathway<br />

to become active may exist, <strong>and</strong> the influence <strong>of</strong> environmental conditions such as pH, temperature, <strong>and</strong> salt must be considered.<br />

INTRODUCTION<br />

b-Sheet-based assemblies have attracted much interest from bonding; <strong>and</strong> 3), reach an alternative ‘‘nonnative’’ global<br />

multidisciplinary researchers because <strong>of</strong> their association free energy minimum (10). Therefore, investigators have<br />

with a variety <strong>of</strong> diseases <strong>and</strong> their emerging potential in emphasized underst<strong>and</strong>ing <strong>and</strong> inhibiting amyloid formation<br />

material science <strong>and</strong> biotechnology (1,2). Protein misfolding more so than amyloid dissociation <strong>and</strong> clearance. Although<br />

<strong>and</strong> self-assembly into highly ordered b-sheet-rich fibrillar the stability <strong>of</strong> b-sheet-rich amyloid fibrils against proteases,<br />

assemblies known as amyloid fibrils are common features acids, <strong>and</strong> chemical denaturants has been shown, increasing<br />

<strong>of</strong> a growing class <strong>of</strong> systemic <strong>and</strong> neurodegenerative evidence from human (11) <strong>and</strong> in vitro studies indicates that<br />

diseases, including Alzheimer’s, Parkinson’s, <strong>and</strong> Hunting- a dynamic structure exists within amyloid fibrils <strong>and</strong><br />

ton’s diseases; senile systemic amyloidoses; type II diabetes suggests that the process <strong>of</strong> amyloid formation is reversible<br />

(3,4); <strong>and</strong> many others. The ability to fibrillate is independent (12). These findings, along with the fact that strategies aimed<br />

<strong>of</strong> the original native structure <strong>of</strong> the protein, whose amino at stabilizing amyloid fibrils <strong>and</strong>/or accelerating their clear-<br />

acid sequence primarily appears to play a key role in terms ance seem to reverse the disease phenotype (13,14), suggest<br />

<strong>of</strong> filament arrangement (5), fibrillation kinetics (6), <strong>and</strong> that a detailed underst<strong>and</strong>ing <strong>of</strong> the formation, stability, <strong>and</strong><br />

overall yield <strong>and</strong> stability <strong>of</strong> the fibrils (7,8). Fibrillation dynamic behavior <strong>of</strong> amyloid fibrils is critically important to<br />

originates under conditions in which proteins are partially the development <strong>of</strong> therapeutic strategies for amyloid<br />

destabilized or completely unfolded (9), <strong>and</strong> the formation diseases. Thus, the reversible untangling <strong>of</strong> amyloid archi-<br />

<strong>of</strong> the amyloid fibrils reflects an alternative to the native tecture <strong>and</strong> intrafibrillar packing <strong>of</strong> the b-pleated sheets is<br />

packing conformational struggle <strong>of</strong> a polypeptide chain to a key issue to consider in designing inhibitors <strong>of</strong> fibrillar<br />

1), reduce its surface-accessible area; 2), saturate hydrogen growth (15), <strong>and</strong> insights into the fibrillar assembly mechanisms<br />

may help elucidate the etiology <strong>of</strong> the ‘‘prion<br />

Submitted September30, 2008, <strong>and</strong> acceptedfor publication December1, 2008.<br />

*Correspondence: pablo.taboada@usc.es<br />

Editor: Ruth Nussinov.<br />

Ó 2009 by the Biophysical Society<br />

diseases’’, provided that the subtle structural differences<br />

underlying the puzzling phenomenon <strong>of</strong> ‘‘prion strains’’<br />

can be understood (16).<br />

0006-3495/09/03/2353/18 $2.00 doi: 10.1016/j.bpj.2008.12.3901<br />

147


2354 Juárez et al.<br />

Because <strong>of</strong> the physiological importance <strong>of</strong> human serum<br />

albumin (HSA) as a carrier protein <strong>and</strong> blood pressure regulator,<br />

<strong>and</strong> its propensity to easily aggregate in vitro, HSA has<br />

become a good model for protein aggregation studies. Moreover,<br />

as the phenomenon <strong>of</strong> protein aggregation appears to<br />

reflect certain generic ‘‘polymeric’’ features <strong>of</strong> proteins<br />

(17), studying mechanisms <strong>of</strong> protein aggregation in model<br />

systems is extremely useful for gaining a better underst<strong>and</strong>ing<br />

<strong>of</strong> the molecular mechanisms <strong>of</strong> disease-associated<br />

amyloidogenesis.<br />

Under physiological conditions, HSA consists <strong>of</strong> 585<br />

amino acids in a single polypeptide chain, with a globular<br />

structure composed <strong>of</strong> three main domains that are loosely<br />

joined together through physical forces <strong>and</strong> six subdomains<br />

that are wrapped by disulfide bonds. The protein contains<br />

17 disulfide bridges <strong>and</strong> one free SH group, which facilitates<br />

dimerization <strong>and</strong> also influences higher-order association.<br />

Native HSA lacks any properties that suggest a predisposition<br />

to form amyloid fibrils, since most <strong>of</strong> its sequence (>60%) is<br />

arranged in an a-helix structure, with subsequent tightening<br />

<strong>of</strong> its structure through intramolecular interactions such as<br />

hydrogen bonds. Therefore, serum albumin aggregation is<br />

promoted under conditions that favor partly destabilized<br />

monomers <strong>and</strong> dimers, such as low pH, high temperature,<br />

<strong>and</strong> the presence <strong>of</strong> chemical denaturants (18). In a recent<br />

report (19), we showed that partially destabilized HSA molecules<br />

form amyloid-like fibrils <strong>and</strong> other types <strong>of</strong> aggregates<br />

under different solution conditions. These fibrils feature the<br />

structural characteristics <strong>of</strong> amyloids: x-ray diffraction<br />

(XRD) patterns, affinity to Congo Red (CR) <strong>and</strong> thi<strong>of</strong>lavin<br />

T (ThT), birefringence, <strong>and</strong> high stability. We now extend<br />

that previous work to shed further light on the kinetics <strong>and</strong><br />

hierarchical assembly <strong>of</strong> HSA fibril formation, linking the<br />

morphological structural transitions <strong>of</strong> aggregated protein<br />

intermediates to conformational events on protein structure<br />

as analyzed by means <strong>of</strong> different biophysical <strong>and</strong> spectroscopic<br />

methods. In this way, we present a systematic investigation<br />

<strong>of</strong> the relationship between protein conformation<br />

<strong>and</strong> the amyloid-like self-assembly pathway for HSA under<br />

different solution conditions. To this end, we incubated the<br />

protein under different thermal <strong>and</strong> solvent conditions <strong>and</strong><br />

analyzed the protein conformation changes upon incubation.<br />

We additionally imaged different structural intermediates on<br />

the HSA fibrillation pathway depending on solution conditions.<br />

In this way, we sought to uncover the structural<br />

features underlying the formation <strong>of</strong> possibly cytotoxic<br />

HSA assemblies.<br />

MATERIALS AND METHODS<br />

Materials<br />

HSA (70024-90-7), CR, <strong>and</strong> ThT were obtained from Sigma (St. Louis,<br />

MO) <strong>and</strong> used as received. All other chemicals were <strong>of</strong> the highest purity<br />

available.<br />

Biophysical Journal 96(6) 2353–2370<br />

Preparation <strong>of</strong> HSA solutions<br />

Protein was used after further purification by liquid chromatography using<br />

a Superdex 75 column equilibrated with 0.01 M phosphate. Experiments<br />

were carried out using double-distilled, deionized, <strong>and</strong> degassed water.<br />

The buffer solutions used were glycine þ HCl (I ¼ 0.01 M) for pH 3.0,<br />

<strong>and</strong> sodium monophosphate-sodium diphosphate for pH 7.4 (I ¼ 0.01 M),<br />

respectively. HSA was dissolved in each buffer solution to a final concentration<br />

<strong>of</strong> typically 20 mg/mL <strong>and</strong> dialyzed extensively against the proper<br />

buffer. Protein concentration was determined spectrophotometrically using<br />

a molar absorption coefficient <strong>of</strong> 35,219 M 1 cm 1 at 280 nm (20). Before<br />

incubation, the solution was filtered through a 0.2 mm filter into sterile test<br />

tubes. Samples were incubated at a specified temperature in a refluxed<br />

reactor. Samples were taken out at intervals <strong>and</strong> stored on ice before addition<br />

<strong>of</strong> CR or ThT.<br />

Seeding solutions<br />

To test whether seeding with preformed aggregates increases the rate <strong>of</strong><br />

HSA aggregation under the different conditions in which fibrils are formed,<br />

a protein solution was incubated for 24 h <strong>and</strong> an aliquot that corresponded to<br />

10% (w/w) <strong>of</strong> the total protein concentration was then added to a fresh<br />

protein solution.<br />

CR binding<br />

Changes in the absorbance <strong>of</strong> CR dye produced by binding onto HSA were<br />

measured in an ultraviolet-visible spectrophotometer (DU series 640; Beckman<br />

Coulter, Fullerton, CA) operating at 190–1100 nm. All measurements<br />

were made in the wavelength range <strong>of</strong> 220–500 nm in matched quartz<br />

cuvettes. Protein solutions were diluted 20- to 200-fold into a buffer solution<br />

with 5 mM <strong>of</strong> CR (Acros Organics, Geel, Belgium). Spectra in the presence<br />

<strong>of</strong> the dye were compared with those <strong>of</strong> the buffer containing CR in the<br />

absence <strong>of</strong> protein <strong>and</strong> also with those corresponding to the protein solution<br />

without dye.<br />

ThT spectroscopy<br />

Protein <strong>and</strong> ThT were dissolved in the proper buffer at a final protein/dye<br />

molar ratio <strong>of</strong> 50:1. Samples were continuously stirred during measurements.<br />

Fluorescence was measured in a Cary Eclipse fluorescence spectrophotometer<br />

equipped with a temperature control device <strong>and</strong> a multicell sample holder<br />

(Varian Instruments, Palo Alto, CA). Excitation <strong>and</strong> emission wavelengths<br />

were 450 <strong>and</strong> 482 nm, respectively. All intensities were background-corrected<br />

for the ThT fluorescence in the respective solvent without the protein.<br />

Protein fluorescence<br />

To examine the conformational variations around the Tryp residue <strong>of</strong> HSA,<br />

fluorescence emission spectra were recorded with a Cary Eclipse fluorescence<br />

spectrophotometer equipped with a temperature control device <strong>and</strong><br />

a multicell sample holder (Varian Instruments). HSA samples were excited<br />

at 295 nm, which provides no excitation <strong>of</strong> tyrosine residues <strong>and</strong>, therefore<br />

does not cause emission or energy transfer to the lone side chain. Slit widths<br />

were typically 5 nm.<br />

Circular dichroism<br />

Far- <strong>and</strong> near-ultraviolet (UV) circular dichroism (CD) spectra were<br />

obtained using a JASCO-715 automatic recording spectropolarimeter (Jasco,<br />

Tokyo, Japan) with a JASCO PTC-343 Peltier-type thermostated cell holder.<br />

Quartz cuvettes with 0.2 cm pathlength were used. CD spectra were obtained<br />

from aliquots withdrawn from the aggregation mixtures at the indicated<br />

conditions <strong>and</strong> recorded between 195 <strong>and</strong> 300 nm at 25 C. The mean residue<br />

ellipticity q (deg cm 2 dmol 1 ) was calculated from the formula:<br />

148


Fibrillation Pathway <strong>of</strong> HSA 2355<br />

q ¼ðqobs=10ÞðMRM=lcÞ, where qobs is the observed ellipticity in deg, MRM<br />

is the mean residue molecular mass, l is the optical pathlength<br />

(in centimeters), <strong>and</strong> c is the protein concentration (in g mL 1 ). To calculate<br />

the composition <strong>of</strong> the secondary structure <strong>of</strong> the protein, SELCON3, CON-<br />

TIN, <strong>and</strong> DSST programs were used to analyze far-UV CD spectra. Final<br />

results were assumed when data generated from all programs showed<br />

convergence (21).<br />

XRD<br />

XRD experiments were carried out using a Siemens D5005 rotating anode<br />

x-ray generator. Twin Göbel mirrors were used to produce a well-collimated<br />

beam <strong>of</strong> CuKa radiation (l ¼ 1.5418 A). Samples were put into capillary<br />

with a diameter <strong>of</strong> 0.5 mm. X-ray diffraction patterns were recorded with<br />

an imaging plate detector (AXS F.Nr. J2-394).<br />

Fourier transform infrared spectroscopy<br />

Fourier transform infrared (FTIR) spectra <strong>of</strong> HSA in aqueous solutions were<br />

determined by using an FTIR spectrometer (model IFS-66v; Bruker) equipped<br />

with a horizontal ZnS ATR accessory. The spectra were obtained at<br />

a resolution <strong>of</strong> 2 cm 1 <strong>and</strong> generally 200 scans were accumulated to obtain<br />

a reasonable signal/noise ratio. Solvent spectra were also examined under<br />

the same accessory <strong>and</strong> instrument conditions. Each different sample spectrum<br />

was obtained by digitally subtracting the solvent spectrum from the<br />

corresponding sample spectrum. Each sample solution was repeated three<br />

times to ensure reproducibility <strong>and</strong> averaged to produce a single spectrum.<br />

Transmission electron microscopy<br />

For transmission electron microscopy (TEM), suspensions <strong>of</strong> HSA were<br />

applied to carbon-coated copper grids, blotted, washed, negatively stained<br />

with 2% (w/v) <strong>of</strong> phosphotungstic acid, air dried, <strong>and</strong> then examined with<br />

a Phillips CM-12 transmission electron microscope operating at an accelerating<br />

voltage <strong>of</strong> 120 kV. Samples were diluted 20- to 200-fold when necessary<br />

before deposition on the grids.<br />

Scanning electron microscopy<br />

Suspensions <strong>of</strong> HSA were applied to glass-coated stainless-steel grids,<br />

blotted, washed, air dried, <strong>and</strong> then examined with an LEO-435VP scanning<br />

electron microscope (Leica Microsystems GmbH, Wetlar, Germany) operating<br />

at an accelerating voltage <strong>of</strong> 30 kV. Samples were diluted 20- to<br />

200-fold when necessary before deposition on the grids. Microanalysis <strong>of</strong><br />

the scanning electron microscopy (SEM) samples was also performed to<br />

avoid the presence <strong>of</strong> impurities.<br />

Atomic force microscopy<br />

Atomic force microscopy (AFM) images were recorded in tapping mode<br />

by using a multimode SPM microscope equipped with a Nanoscope IIIa<br />

controller from Digital Instruments (Santa Barbara, CA). The microscope<br />

was coupled to an AS-12 resp. E-scanner <strong>and</strong> an Extender Electronics Module<br />

EX-II, which allows acquisition <strong>of</strong> phase images. The AFM probes were typically<br />

silicon SPM sensors (NCHR Nanosensors, Neuchatel, Switzerl<strong>and</strong>).<br />

Immediately after incubation, the protein samples were diluted 20–400 times<br />

onto freshly cleaved muscovite mica (Sigma) attached to a magnetic steel disc<br />

that served as the sample holder. The abrupt dilution <strong>of</strong> the samples immediately<br />

quenched the concentration-dependent aggregation process. The AFM<br />

samples were dried on air or under nitrogen flow when required. Control<br />

samples (freshly cleaved mica, <strong>and</strong> mica <strong>and</strong> buffer solution) were also investigated<br />

with AFM to exclude possible artifacts. Height <strong>and</strong> phase-shift data<br />

were collected in the trace <strong>and</strong> the respective retrace direction <strong>of</strong> the raster.<br />

The scan rate was tuned proportionally to the area scanned <strong>and</strong> was kept<br />

within the 0.35–2 Hz range.<br />

RESULTS<br />

It is generally accepted that amyloid formation usually is<br />

a result <strong>of</strong> misfolded <strong>and</strong> partially unfolded states acting in<br />

competition with the normal folding pathways (22–24).<br />

The HSA molecule is known to undergo several well-organized<br />

changes during its conformation, usually under nonphysiological<br />

conditions, as follows:<br />

1. The N-F transition between pH 5.0 <strong>and</strong> 3.5 involves the<br />

unfolding <strong>and</strong> separation <strong>of</strong> domain III without significantly<br />

affecting the rest <strong>of</strong> the protein molecule<br />

(25,26). The F form is characterized by a dramatic<br />

increase in viscosity, lower solubility, <strong>and</strong> a significant<br />

loss <strong>of</strong> helical content.<br />

2. The F-E transition occurs between pH 3.5 <strong>and</strong> 1.2, <strong>and</strong> is<br />

accompanied by a further protein expansion with a loss<br />

<strong>of</strong> the intradomain helices <strong>of</strong> domain I. In addition,<br />

the E form involves an increase in protein intrinsic<br />

viscosity <strong>and</strong> a rise in the hydrodynamic axial ratio from<br />

~4 to 9 (27).<br />

3. The N-B transition occurs between pH 7.0 <strong>and</strong> 9.0, with<br />

a slight reduction in helical content affecting the two interdomain<br />

helices <strong>and</strong> a small increase in sheet structure (28).<br />

4. In the presence <strong>of</strong> denaturant agents, such as urea, HSA<br />

shows a two-step, three-state transition with an intermediate<br />

(I) characterized by unfolding <strong>of</strong> the domain III <strong>and</strong><br />

partial but significant loss <strong>of</strong> native conformation <strong>of</strong><br />

domain I (29).<br />

Therefore, an easy way to obtain at least partially denaturated<br />

states is to induce a temperature-induced or solventinduced<br />

protein denaturation process through incubation.<br />

Thus, HSA was subjected to conditions previously found to<br />

be effective for protein aggregation, in particular those that<br />

induce amyloid-like fibril formation (19,30). HSA solutions<br />

were incubated at a concentration <strong>of</strong> 20 mg/mL in 0.01 M<br />

sodium phosphate buffer, pH 7.4, or 0.01 M glycine buffer,<br />

pH 3.0, in the presence <strong>of</strong> 0 or 50 mM <strong>of</strong> added NaCl at<br />

25 Cor65 C for 15 days. We chose 65 C as the incubation<br />

temperature because HSA temperature-induced denaturation<br />

takes place through a two-state transition with a first melting<br />

temperature, Tm, <strong>of</strong> ~56 C <strong>and</strong> a second Tm <strong>of</strong> ~62 C(31,32)<br />

as a consequence <strong>of</strong> the sequential unfolding <strong>of</strong> the different<br />

domains <strong>of</strong> the protein, in particular, the IIA <strong>and</strong> IIIA<br />

subdomains. Moreover, as the pH becomes more acidic,<br />

T m becomes lower. We confirmed these melting temperatures<br />

under our solution conditions by fluorescence spectroscopy,<br />

<strong>and</strong> the results were in close agreement with previously<br />

reported data (see Table S1 in the Supporting Material).<br />

Therefore, it can be inferred that aggregation is unfavorable<br />

below 65 C because protein folding competes with <strong>and</strong><br />

suppresses amyloid formation. It is known that hydrogen<br />

bonding is weakened as temperature rises, but the hydrophobic<br />

interaction becomes strengthened (33). As the heating<br />

process continues, some <strong>of</strong> the cooperative hydrogen bonds<br />

Biophysical Journal 96(6) 2353–2370<br />

149


2356 Juárez et al.<br />

that stabilized helical structure begin to break <strong>and</strong> expose<br />

hydrophobic groups to the solvent, partially unfolding the<br />

protein structure, which favors aggregation.<br />

Physiological conditions: kinetics <strong>and</strong> amyloid<br />

self-assembly <strong>of</strong> HSA<br />

The propensity <strong>of</strong> protein solutions to form amyloid-like<br />

aggregates under physiological conditions was assessed by<br />

means <strong>of</strong> ThT fluorescence <strong>and</strong> CR absorbance measurements.<br />

These two dyes specifically bind to ordered b-sheet<br />

aggregates <strong>and</strong>, notably, to amyloid fibrils (34,35). Both<br />

assays are necessary because positive ThT binding sometimes<br />

does not occur with certain amyloid fibril systems (36).<br />

Before incubation, native HSA does not display a capacity<br />

to bind ThT. When incubated at physiological pH <strong>and</strong> room<br />

temperature, the fluorescence emission intensity for the<br />

protein at 482 nm is negligible, suggesting that HSA does<br />

not form amyloid fibrils or other types <strong>of</strong> fluorescent aggregates<br />

under these conditions. When the temperature is raised<br />

to 65 C, a time-dependent increase in fluorescence is<br />

observed (Fig. 1, a <strong>and</strong> b). The kinetics <strong>of</strong> HSA involved<br />

the continuous rising <strong>of</strong> ThT fluorescence during the early<br />

periods <strong>of</strong> the incubation procedure, <strong>and</strong> exhibited no discernible<br />

lag phase until a quasi-plateau region was attained in the<br />

FIGURE 1 Time evolution <strong>of</strong> ThT fluorescence in HSA solutions incubated<br />

at 65 C at pH 7.4 in the (a) absence <strong>and</strong> (b) presence <strong>of</strong> 50 mM NaCl.<br />

Biophysical Journal 96(6) 2353–2370<br />

timescale analyzed. The addition <strong>of</strong> electrolyte favors a faster<br />

formation <strong>of</strong> amyloid fibrils as a consequence <strong>of</strong> the screening<br />

<strong>of</strong> electrostatic repulsions between protein molecules (37). In<br />

fact, a gel phase can be observed after 4 days <strong>of</strong> incubation,<br />

which denotes an enhanced development <strong>of</strong> fibril formation<br />

because cross-links can be formed more easily, resulting in<br />

a lower critical percolation concentration.<br />

ThT fluorescence curves were fitted by means <strong>of</strong> nonlinear<br />

square curve-fitting to a stretched exponential function<br />

F ¼ F N þ DFexpð ½ksptŠ n Þ to obtain information on the<br />

kinetics <strong>of</strong> amyloid formation. F, F N , <strong>and</strong> DF are the<br />

observed fluorescence intensity at time t, the final fluorescence<br />

intensity, <strong>and</strong> the fluorescence amplitude, respectively,<br />

<strong>and</strong> k sp is the rate <strong>of</strong> spontaneous fibril formation. Although<br />

the interpretation <strong>of</strong> the parameters involved in this equation<br />

is not straightforward, these are useful for empiric descriptions<br />

<strong>of</strong> the complex reactions whose kinetics is not easily<br />

modeled (38,39). Values <strong>of</strong> ksp, n, <strong>and</strong> DF determined in<br />

this way under the different conditions are shown in Table 1.<br />

Values <strong>of</strong> n < 1 indicated that the kinetics can be approximated<br />

to several exponential functions indicative <strong>of</strong> the<br />

existence <strong>of</strong> multiple events in the amyloid formation.<br />

Furthermore, larger ksp values in the presence <strong>of</strong> electrolyte<br />

corroborate the more-efficient aggregation due to electrostatic<br />

screening under high ionic strength conditions.<br />

CR absorption also corroborates the formation <strong>of</strong> amyloid<br />

fibrils, displaying a progressive red shift from 495 to ~530 nm<br />

<strong>of</strong> the differential absorption maximum at 65 C in both the<br />

absence <strong>and</strong> presence <strong>of</strong> added electrolyte, a typical feature<br />

<strong>of</strong> amyloid fibers (35). Since absorption <strong>of</strong> HSA alone after<br />

incubation contributes to only ~15% <strong>of</strong> the increase in CR<br />

absorption, the change in absorption is mainly caused by<br />

the formation <strong>of</strong> CR-binding species (see Fig. S1 in the Supporting<br />

Material).<br />

Fibrillation is independent <strong>of</strong> seeding<br />

at physiological pH<br />

Typical fibrillation processes involve a lag phase followed by<br />

a relatively rapid elongation phase that stabilizes when all<br />

TABLE 1 Kinetic parameters <strong>of</strong> the self-assembly process <strong>of</strong><br />

HSA solutions<br />

F N<br />

DF n ksp (h 1 )<br />

pH 7.4<br />

65 C<br />

0 mM NaCl 85 83 0.56 0.019<br />

50 mM NaCl 85 83 0.82 0.064<br />

pH 3.0<br />

25 C<br />

0 mM NaCl 10 8 1.34 0.007<br />

50 mM Na Cl 13 11 1.57 0.007<br />

65 C<br />

0 mM NaCl 8/30 6/22 1.25/3.6 0.144/0.005<br />

50 mM NaCl 59 57 0.96 0.010<br />

150


Fibrillation Pathway <strong>of</strong> HSA 2357<br />

monomers have been incorporated into fibrils (40–42). In our<br />

case, the absence <strong>of</strong> a lag phase at pH 7.4 would suggest that<br />

nuclei are either formed very rapidly or the aggregation<br />

process we monitor is not a classical nucleation-based fibrillation.<br />

If nuclei consist <strong>of</strong> more than one molecule, their<br />

formation will be reduced if the HSA concentration is lowered.<br />

Nevertheless, when the protein concentration was<br />

decreased to 0.5 <strong>and</strong> 2 mg/mL at 65 C, we did not observe<br />

the appearance <strong>of</strong> a lag phase (figure not shown). In addition,<br />

another feature that confirms a continuous fibrillation process<br />

without a nucleation step is the absence <strong>of</strong> any remarkable<br />

effect on the fluorescence curves when protein seeds were<br />

added to protein solutions followed by subsequent incubation<br />

(see Fig. S2). The aggregation rate during the growth phase<br />

was unchanged by the addition <strong>of</strong> preformed aggregates <strong>and</strong><br />

followed apparent first-order kinetics. These characteristics<br />

suggest that aggregation does not require nucleation, i.e.,<br />

each protein monomer association step is bimolecular <strong>and</strong><br />

effectively irreversible, <strong>and</strong> there is no energy barrier to aggregate<br />

growth. As discussed in detail below, TEM pictures<br />

showed the formation <strong>of</strong> spherical oligomers after very short<br />

incubation times. This usually occurs by a mechanism <strong>of</strong> classical<br />

coagulation, or downhill polymerization (43), that does<br />

not require a nucleation step.<br />

Structural changes upon aggregation<br />

at physiological conditions: secondary structure<br />

To gain insight into the structural protein modifications upon<br />

formation <strong>of</strong> amyloid-like aggregates, we recorded CD, FTIR,<br />

<strong>and</strong> tryptophan (Tryp) fluorescence spectra. As a supplement<br />

to ThT <strong>and</strong> CR assays, which provide information only about<br />

the formation <strong>of</strong> fibrillar protein assemblies, far-UV CD <strong>and</strong><br />

FTIR data reveal the overall protein secondary structural<br />

composition <strong>and</strong> their evolution to form amyloid-like or amorphous<br />

aggregates. On the other h<strong>and</strong>, near-UV CD <strong>and</strong> Tryp<br />

fluorescence data denote changes in protein tertiary structure<br />

(for the latter technique, particularly in domain II <strong>of</strong> HSA).<br />

Fig. 2, a <strong>and</strong> b, show far-UV CD spectra <strong>of</strong> HSA at pH 7.4<br />

at 25 C <strong>and</strong> 65 C. The spectra at room temperature in the<br />

absence <strong>of</strong> added salt show two minima—one at 208 <strong>and</strong><br />

other at 222 nm—characteristic <strong>of</strong> helical structure, which<br />

remains unmodified upon incubation. When electrolyte is<br />

present, a small decrease in ellipticity, [q], occurs as a consequence<br />

<strong>of</strong> small changes in protein structure originating from<br />

the formation <strong>of</strong> some amorphous aggregates in solution, as<br />

shown in Fig. 2 a. In contrast, when the temperature is raised<br />

to 65 C, the minimum at 222 nm progressively disappears<br />

<strong>and</strong> [q] at 208 nm also strongly reduces (Fig. 2 b). This indicates<br />

that high-temperature conditions spawn intermediates<br />

that are clearly less helical than the starting conformations.<br />

This change in CD spectra suggests the increment <strong>of</strong> either<br />

b-sheet or loop structures, as discussed in detail further<br />

below. The intensity loss <strong>of</strong> the 222 nm b<strong>and</strong> indicates that<br />

the increase in r<strong>and</strong>om coil/b-sheet conformations is accom-<br />

FIGURE 2 (a) Far-UV spectra <strong>of</strong> HSA solutions at 25 C in the presence<br />

<strong>of</strong> 50 mM NaCl at pH 7.4 at 1), 0 h; 2), 24 h; 3), 48 h; 4), 100 h; 5), 200 6),<br />

<strong>and</strong> 7), 250 h <strong>of</strong> incubation. (b) Far-UV spectra <strong>of</strong> HSA solutions at 65 Cin<br />

the presence <strong>of</strong> 50 mM NaCl at 1), 0 h; 2), 12 h; 3), 24 h; <strong>and</strong> 4), 48 h <strong>of</strong><br />

incubation.<br />

panied by a reduction in the a-helical content <strong>of</strong> the protein<br />

structure. At longer incubation times in the absence <strong>of</strong> added<br />

salt in solution (~72 h), the CD spectrum resembles that<br />

typical <strong>of</strong> proteins with high proportions <strong>of</strong> b-sheet structure,<br />

which are characterized by a minimum at ~215–220 nm <strong>and</strong><br />

small [q] values as a consequence <strong>of</strong> the presence <strong>of</strong><br />

amyloid-like aggregates in solution. On the other h<strong>and</strong>, no<br />

significant changes in far-UV CD spectra are observed<br />

when electrolyte is added to solutions, except that the characteristic<br />

features <strong>of</strong> b-sheet structure are present at earlier incubation<br />

times at elevated temperature. A larger decrease in [q]<br />

is also detected as a consequence <strong>of</strong> the increased number <strong>and</strong><br />

size <strong>of</strong> scattering objects in solution, which agrees with the<br />

formation <strong>of</strong> a fibrillar gel upon longer incubation times.<br />

Fig. 3 depicts the temporal evolution <strong>of</strong> the secondary<br />

structure composition as revealed by CD analysis (21). At<br />

pH 7.4 <strong>and</strong> room temperature, the initial a-helix content<br />

is ~59%, the b-sheet conformation is ~5%, the turn content<br />

Biophysical Journal 96(6) 2353–2370<br />

151


2358 Juárez et al.<br />

is ~13%, <strong>and</strong> the remaining r<strong>and</strong>om coil content is ~23%, in<br />

agreement with previous reports (25). In the presence <strong>of</strong><br />

added electrolyte, no significant alterations in the structure<br />

compositions are observed upon incubation; the initial<br />

a-helix content is slightly reduced to ~55%, in contrast to<br />

a very small increase in turn <strong>and</strong> b-sheet conformations.<br />

After ~150 h <strong>of</strong> incubation, an additional slight decrease in<br />

a-helix is observed due to an enhancement <strong>of</strong> protein aggregation,<br />

i.e., the formation <strong>of</strong> a certain amount <strong>of</strong> protein<br />

aggregates is observed, which is a characteristic feature <strong>of</strong><br />

globular proteins under high ionic strength conditions.<br />

On the other h<strong>and</strong>, protein conformational compositions at<br />

65 C after incubation indicate an increase in b-sheet conformation<br />

from ~5% to ~21% at the expense <strong>of</strong> the a-helix<br />

content (which diminishes from ~59% to ~20%) in the<br />

absence <strong>of</strong> electrolyte. This is also reflected by the decrease<br />

in the CD signal <strong>and</strong> the shift <strong>of</strong> the spectral minimum from<br />

208 nm to longer wavelengths, as noted above. Coil <strong>and</strong> turn<br />

conformations also change through the incubation process,<br />

with values <strong>of</strong> ~27–34% <strong>and</strong> ~17–21%, respectively. The<br />

alteration in structure composition seems to be stronger<br />

when salt is added, due to electrostatic screening, which<br />

favors protein disruption <strong>and</strong> association by modifying the<br />

balance <strong>of</strong> interactions, with final b-sheet <strong>and</strong> a-helix<br />

contents <strong>of</strong> ~26% <strong>and</strong> ~19%, respectively (Fig. 3). Moreover,<br />

changes in secondary structure take place during the<br />

first part <strong>of</strong> the incubation process in both the absence<br />

(~90 h) <strong>and</strong> presence <strong>of</strong> electrolyte (~50 h), respectively,<br />

<strong>and</strong> occur more rapidly under the latter condition. At very<br />

long incubation times (>150 h) under high ionic strength<br />

conditions, the increased scattering from the fibrillar aggregates<br />

makes it difficult to estimate the secondary structure;<br />

thus, these results are not shown in Fig. 3.<br />

FTIR spectra were recorded at the beginning <strong>and</strong> end <strong>of</strong><br />

the incubation process by monitoring the observed changes<br />

Biophysical Journal 96(6) 2353–2370<br />

FIGURE 3 Time evolution <strong>of</strong> secondary structure compositions<br />

<strong>of</strong> HSA solutions at pH 7.4 at 25 C(a <strong>and</strong> b)or65 C<br />

(c <strong>and</strong> d) in the absence <strong>and</strong> presence <strong>of</strong> 50 mM NaCl, respectively.<br />

(:) a-helix, (-) b-turn, (C) unordered, <strong>and</strong> (;)<br />

b-sheet conformations.<br />

in the shape <strong>and</strong> frequency <strong>of</strong> the amide I <strong>and</strong> II b<strong>and</strong>s, <strong>and</strong><br />

the results corroborated the CD data. Before incubation, two<br />

major absorption peaks in the spectral region <strong>of</strong> interest were<br />

observed: the amide I b<strong>and</strong> at 1653 (1652) cm 1 <strong>and</strong> the amide<br />

II b<strong>and</strong> at 1542 (1544) cm 1 in both the original <strong>and</strong> second<br />

derivative IR spectra, respectively. This indicates the predominant<br />

structural contribution <strong>of</strong> major a-helix <strong>and</strong> minor<br />

r<strong>and</strong>om coil structures, in agreement with CD data (44–46).<br />

For the amide I b<strong>and</strong> (Fig. 4 a), a shoulder at ~1630 cm 1<br />

can be also observed in the second derivative spectra that is<br />

related to intramolecular b-sheet structure. Additional peaks<br />

at ~1689 <strong>and</strong> 1514 cm 1 would correspond to b-turn <strong>and</strong> tyrosine<br />

absorption, respectively (45). All <strong>of</strong> these peaks were also<br />

observed in the presence <strong>of</strong> electrolyte at room temperature as<br />

FIGURE 4 Second derivative <strong>of</strong> FTIR spectra at pH 7.4 <strong>of</strong> (a) native HSA<br />

at 25 C before incubation, (b) HSA at 25 C in the presence <strong>of</strong> 50 mM NaCl,<br />

<strong>and</strong> (c) HSA at 65 C in the absence <strong>of</strong> electrolyte.<br />

152


Fibrillation Pathway <strong>of</strong> HSA 2359<br />

incubation proceeded (Fig. 4 b), compatible with little changes<br />

in secondary structure composition, as noted above.<br />

After incubation at 65 C, a red shift <strong>of</strong> the amide I b<strong>and</strong><br />

from 1652 to 1658 cm 1 <strong>and</strong> a blue shift <strong>of</strong> the amide II<br />

b<strong>and</strong> to 1542 cm 1 in the second derivative spectrum are<br />

indicative <strong>of</strong> a certain increase <strong>of</strong> disordered structure, as<br />

revealed by far-UV CD (Fig. 4 c). The appearance <strong>of</strong><br />

a well-defined peak around 1628 cm 1 (1626 cm 1 in the<br />

presence <strong>of</strong> added salt) points to a structural transformation<br />

from an intramolecular hydrogen-bonded b-sheet to an intermolecular<br />

hydrogen-bonded-b-sheet structure (47). The spectrum<br />

also shows a high-frequency component (~1692 cm 1 )<br />

that would suggest the presence <strong>of</strong> an antiparallel b-sheet<br />

(48). In addition, a little shoulder around 1534 cm 1 also<br />

was assigned to b-sheet after incubation at high temperature<br />

(49). The increase in the b<strong>and</strong> associated with b-sheet in the<br />

FTIR spectra correlates well with the large changes in CD<br />

spectra. On the other h<strong>and</strong>, peak shifts are slightly more abrupt<br />

when electrolyte is present in solution because the aggregation<br />

is stronger (plot not shown). This leads to the formation<br />

<strong>of</strong> a fibrillar gel favored by the decrease in electrostatic repulsions<br />

<strong>and</strong> a change in the hydration layer surrounding the<br />

protein molecules that allows interfibrillar attachment.<br />

Structural changes upon aggregation at<br />

physiological conditions: tertiary structure<br />

At room temperature, the near-UV CD spectrum showed two<br />

minima at 262 <strong>and</strong> 268 nm <strong>and</strong> two shoulders around 275 <strong>and</strong><br />

285 nm, characteristic <strong>of</strong> disulphide <strong>and</strong> aromatic chromophores<br />

<strong>and</strong> the asymmetric environment <strong>of</strong> the latter (Fig. 5<br />

a)(50). These features are significantly retained during incubation<br />

in the presence <strong>of</strong> electrolyte excess despite the formation<br />

<strong>of</strong> some amorphous aggregates in solution, as noted<br />

above (figure not shown). On the other h<strong>and</strong>, when the incubation<br />

temperature is raised to 65 C, important alterations in<br />

the near-UV CD spectra occur in both the absence <strong>and</strong> presence<br />

<strong>of</strong> electrolyte: ellipticity decreases, <strong>and</strong> the minima at<br />

262 <strong>and</strong> 268 nm progressively disappear as incubation<br />

proceeds. In addition, a significant loss <strong>of</strong> fine structure<br />

detectable in the region <strong>of</strong> 270–295 nm also occurs. Nevertheless,<br />

these changes in near-UV CD data are less marked<br />

than in far-UV measurements because secondary structural<br />

changes are more sensitive to temperature than tertiary structural<br />

changes (51). These effects corroborate alterations in<br />

tertiary structure upon fibrillation conditions.<br />

The behavior described by near-CD UV measurements is<br />

also supported by tryptophanyl fluorescence data. HSA has<br />

a single tryptophanyl residue, Tryp 214 , located in domain<br />

II. A very small change in fluorescence emission occurs<br />

with incubation at room temperature in the presence <strong>of</strong> electrolyte,<br />

confirming the existence <strong>of</strong> little variations in tertiary<br />

structure, in agreement with previous data (not shown). In<br />

contrast, when the temperature is raised to 65 C, a 4 nm<br />

hypsochromic shift (from 341 to 337 nm) takes place accom-<br />

FIGURE 5 (a) Near-UV CD spectra <strong>of</strong> HSA solutions at pH 7.4 <strong>and</strong> 65 C<br />

in the absence <strong>of</strong> electrolyte at 1), 0 h; 2), 6 h; 3), 12 h; 4), 24 h; <strong>and</strong> 5), 48 h.<br />

(b) Time evolution <strong>of</strong> Trp fluorescence <strong>of</strong> HSA solutions at 65 C in the (-)<br />

absence <strong>and</strong> (B) presence <strong>of</strong> 50 mM NaCl.<br />

panied by a reduction in fluorescence intensity: the domain II<br />

<strong>of</strong> HSA unfolds in such a way that the Trp 214 residue <strong>of</strong> HSA<br />

located at the bottom <strong>of</strong> a 12 A˚ deep crevice (52) finds itself<br />

in a more hydrophobic environment (32). As incubation at<br />

elevated temperature proceeds, conversion <strong>of</strong> protein molecules<br />

to fibrils is accompanied by an additional burial <strong>of</strong><br />

the Tryp residue, as indicated by the decrease <strong>of</strong> the Tryp<br />

fluorescence emission intensity (Fig. 5 b) <strong>and</strong> the further<br />

blue shift <strong>of</strong> the emission maximum from 337 to 333 nm<br />

(53). All <strong>of</strong> this points to the strong involvement <strong>of</strong> tertiary<br />

structure changes <strong>of</strong> domain II <strong>of</strong> HSA in the fibrillation<br />

mechanism. This change is enhanced <strong>and</strong> occurs more<br />

rapidly in the presence <strong>of</strong> added electrolyte, which confirms<br />

a larger structure alteration as a consequence <strong>of</strong> enhanced<br />

intermolecular interactions to give fibrillar assemblies. This<br />

also agrees with the increasing content <strong>of</strong> b-sheet conformation<br />

revealed by CD <strong>and</strong> FTIR measurements.<br />

In situ observation <strong>of</strong> fibrillar structures at<br />

physiological pH: TEM<br />

TEM pictures were recorded at different stages <strong>of</strong> the incubation<br />

period for each <strong>of</strong> the conditions described above.<br />

Distinct time-dependent morphological stages can be<br />

Biophysical Journal 96(6) 2353–2370<br />

153


2360 Juárez et al.<br />

observed in these images. Thus, at room temperature neither<br />

fibrils nor other types <strong>of</strong> aggregates are detected, except for<br />

small amorphous protein clusters observed in the presence <strong>of</strong><br />

electrolyte after a long incubation period (150 h), which<br />

possess a largely helical structure (Fig. 6 a). In contrast,<br />

when the temperature is raised to 65 C, fibril formation is<br />

observed. Electron microscopy indicates that aggregation<br />

leads first to a globular species that subsequently converts<br />

to fibrils with a curly morphology. The fibrillation pathway<br />

in the presence <strong>of</strong> electrolyte is very similar to that observed<br />

in its absence but it takes place in a shorter timescale, in<br />

agreement with previous results. Fig. 6, b–j, show electron<br />

micrographs <strong>of</strong> the sample heated at 65 C at different steps<br />

<strong>of</strong> the incubation period in the presence <strong>of</strong> 50 mM <strong>of</strong> electrolyte.<br />

The number <strong>and</strong> length <strong>of</strong> the fibrils has increased in<br />

relation to other structures, although several morphologies<br />

can be observed throughout incubation.<br />

Small spherical clusters <strong>of</strong> ~20 nm formed by protein oligomers<br />

are observed (Fig. 6 b) at short incubation times (5 h).<br />

These aggregates present relatively few changes in their<br />

tertiary <strong>and</strong> secondary structures, as shown by CD <strong>and</strong> fluorescence<br />

data. With further incubation (Fig. 6 c, 15 h), a certain<br />

elongation <strong>of</strong> these spherical aggregates can be observed. This<br />

bead-like structure at short incubation times arises from what<br />

appears to be attractive interactions between spherical<br />

a b<br />

c<br />

e f<br />

d-1) d-2)<br />

proteins aggregates, as shown in Fig. 6 d-1 (see also<br />

Fig. S3), which may result in an increased exposure <strong>of</strong> hydrophobic<br />

residues, giving rise to more elongated structures. This<br />

elongation involves a conformational conversion <strong>of</strong> protein<br />

structure to consolidate the structure, <strong>and</strong> in all probability<br />

it implies changes in the hydrogen-bonding status (Fig. 6<br />

d-2). This is in agreement with a further development in<br />

ThT fluorescence <strong>and</strong> decreases in both helical content <strong>and</strong><br />

Tryp fluorescence at this incubation point, as shown previously.<br />

On the other h<strong>and</strong>, we did not find evidence <strong>of</strong><br />

formation <strong>of</strong> elongated structures by longitudinal fusion <strong>of</strong><br />

oligomers, as recently reported (54).<br />

Bead-like structures progressively become more elongated<br />

upon incubation (35 h) due to mutual interactions<br />

between these structures <strong>and</strong> subsequent annealing, <strong>and</strong><br />

convert into short prot<strong>of</strong>ibrils (Fig. 6 e), in agreement with<br />

a decrease in helical structure as revealed by CD. Alterations<br />

in the conformational structure <strong>of</strong> these oligomers <strong>and</strong> subsequent<br />

elongation via monomer addition may also be present;<br />

however, the TEM resolution did not allow us to confirm<br />

that. Several authors reported a tendency for these beadlike<br />

structures to transform into fibrillar structures at elevated<br />

temperatures caused by partial unfolding <strong>of</strong> the protein molecules<br />

<strong>and</strong> giving rise to conditions conductive to fibril formation<br />

(55–57).<br />

g h<br />

i j<br />

FIGURE 6 TEM pictures <strong>of</strong> the different stages <strong>of</strong> the HSA fibrillation process at pH 7.4: (a) at25 C in the presence <strong>of</strong> 50 mM NaCl after 150 h <strong>of</strong> incubation,<br />

<strong>and</strong> at 65 C in the presence <strong>of</strong> 50 mM NaCl after (b) 5 h; (c) <strong>and</strong> (d) 15 h (part d shows the elongation <strong>of</strong> oligomers to give bead-like structures); (e)35h<br />

(where short prot<strong>of</strong>ibrils are observed); (f)45h;(g) 50 h (where long curly fibrils are seen); <strong>and</strong> (h–k) after 72 h. Part i shows the addition <strong>of</strong> oligomers to mature<br />

fibrils, j shows the association <strong>of</strong> mature fibrils in bundles, <strong>and</strong> k shows mature fibrils with ribbon-like structure.<br />

Biophysical Journal 96(6) 2353–2370<br />

k<br />

154


Fibrillation Pathway <strong>of</strong> HSA 2361<br />

Further incubation results in the presence <strong>of</strong> more fibrils,<br />

which increase in length as incubation proceeds (Fig. 6,<br />

f <strong>and</strong> g). After 2 days <strong>of</strong> incubation, numerous longer<br />

curly-branched <strong>and</strong> interconnected fibrils are present<br />

(Fig. 6 g), with lengths <strong>and</strong> widths characteristic <strong>of</strong> classical<br />

amyloid fibrils (i.e., between 0.5 <strong>and</strong> several micrometers in<br />

length <strong>and</strong> 9–10 nm in width (3,5,9)), as detected by TEM<br />

<strong>and</strong> AFM (see Fig. 6 g <strong>and</strong> Fig. S4). The appearance <strong>of</strong><br />

this structure coincides well with the plateau region <strong>of</strong> ThT<br />

binding, the far UV-CD pr<strong>of</strong>iles characterized by minima<br />

around 215–220 nm, <strong>and</strong> the intermolecular b-sheet structure<br />

observed by FTIR. Most fibrils appear curly <strong>and</strong> interconnected,<br />

<strong>and</strong> some <strong>of</strong> them are even circular, as observed<br />

previously for other proteins, such as b 2 microglobulin<br />

(58) <strong>and</strong> a-crystallin (56).<br />

If the incubation time is extended further (3 days), one can<br />

observe mature straight fibrils, which can be seen as a structural<br />

evolution <strong>of</strong> curly fibrils (Fig. 6 h). Mature fibrils are<br />

thicker <strong>and</strong> stiffer than single fibrils <strong>and</strong> seem to be formed<br />

by lateral (side-by-side) assembly <strong>of</strong> two or more individual<br />

filaments. These mature fibrils are less numerous in the<br />

absence <strong>of</strong> electrolyte because the electrostatic screening is<br />

lower, avoiding direct contact between constitutive fibrils.<br />

We have also observed that lateral interactions <strong>of</strong> single<br />

particles collaborate in growing these straight mature fibrils<br />

(57). Scrutiny <strong>of</strong> the protein aggregates indicates that HSA<br />

particles <strong>and</strong> clusters tend to align within the immediate<br />

vicinity <strong>of</strong> the fibers (Fig. 6 i), serving the single fiber as<br />

a lateral template or scaffold for small protein molecules,<br />

<strong>and</strong> would constitute a subcomponent in mature fibrils. It<br />

is worth mentioning that some fibril solutions when analyzed<br />

by SEM show the existence <strong>of</strong> very long fibers, exceeding<br />

200 mm (see Fig. S5). We think that this effect may result<br />

from the air-drying process favoring attractive interactions<br />

between single fibers during the solvent removal.<br />

In addition, a structural diversity <strong>of</strong> mature fibrils is noted:<br />

flat ribbons are observed in solution, as well as long, straight<br />

fibril ensembles (Fig. 6 j). This polymorphism may arise<br />

from variations in the quaternary structure, the manner in<br />

which prot<strong>of</strong>ilaments self-associate, or the prot<strong>of</strong>ilament<br />

substructure (e.g., in the details <strong>of</strong> hydrogen-bonding<br />

networks <strong>and</strong> side-chain packing) (59). Finally, one <strong>of</strong> the<br />

characteristic traits <strong>of</strong> the mature amyloid fibrils is their<br />

tendency to bend, twist, <strong>and</strong> agglomerate. Fig. 6 k shows<br />

laterally connected fibers that split over a certain distance<br />

or overlap each other. The extent <strong>and</strong> rate <strong>of</strong> this growth is<br />

dependent on solution conditions <strong>and</strong> lateral interactions<br />

between fibrils, which are responsible for the thickening <strong>of</strong><br />

the mature fibrils <strong>and</strong> the formation <strong>of</strong> suprafibrillar aggregates.<br />

Thus, we conclude that pseudo-globular aggregates<br />

rearrange slowly to form linear, curly fibrils. These may be<br />

sufficient to produce a high-affinity template that is subsequently<br />

elongated by monomeric units or other fibrils, <strong>and</strong><br />

can lead to the formation <strong>of</strong> ordered, straight, or ribbonlike<br />

fibrillar structures.<br />

Fibril structure: XRD<br />

The amyloid-like character <strong>of</strong> the fibrillar aggregates detected<br />

by TEM was confirmed by XRD. The XRD image <strong>of</strong> the HSA<br />

fibrils is shown in Fig. 7. Two strong reflections can be<br />

observed: a dominant sharp <strong>and</strong> intense reflection occurs at<br />

4.8 A˚ , <strong>and</strong> one weaker, more diffuse, but still intense reflection<br />

is observed at ~11 A˚ . The 4.8 A˚ meridional reflection<br />

arises from the spacing between hydrogen-bonded individual<br />

str<strong>and</strong>s in the b-sheet structure that lie perpendicular to the<br />

fibril axis, <strong>and</strong> the 11 A˚ equatorial reflection corresponds to<br />

the intersheet spacing, with the b-sheets stacked face to face<br />

to form the core structure <strong>of</strong> prot<strong>of</strong>ilaments (60,61). This indicates<br />

that fibrils possess a cross-b structure, one <strong>of</strong> the diagnostic<br />

hallmarks <strong>of</strong> amyloid structures.<br />

Acidic conditions: kinetics <strong>and</strong> amyloid<br />

self-assembly <strong>of</strong> HSA<br />

We next subjected HSA to acidic pH <strong>and</strong> assessed its propensity<br />

to form amyloid fibrils under such conditions in both the<br />

presence <strong>and</strong> absence <strong>of</strong> 50 mM NaCl. Increases in ThT fluorescence<br />

were observed at both 25 C <strong>and</strong> 65 C, although the<br />

increase was quite small at 25 C. This indicates the formation<br />

<strong>of</strong> a small amount <strong>of</strong> additional b-sheet conformation as<br />

a consequence <strong>of</strong> the formation <strong>of</strong> oligomeric aggregates, as<br />

will be shown below (Fig. 8). At 65 C, the increase in<br />

ThT fluorescence in the absence <strong>of</strong> electrolyte is lower than<br />

that obtained at physiological pH <strong>and</strong> extends over a larger<br />

period <strong>of</strong> time. This can be a result, on the one h<strong>and</strong>, <strong>of</strong> a lower<br />

capacity <strong>of</strong> fibrillation under these conditions (see below) or,<br />

on the other, to the presence <strong>of</strong> a lag phase if compared with<br />

solution to which electrolyte is added, as discussed further<br />

below. After a small increase in ThT fluorescence at relatively<br />

short incubation times, a plateau occurs at ~24–100 h, after<br />

which the ThT fluorescence starts to increase again. Thus, it<br />

FIGURE 7 XRD pattern <strong>of</strong> HSA fibrils.<br />

Biophysical Journal 96(6) 2353–2370<br />

155


2362 Juárez et al.<br />

is thought that under acidic conditions in the absence <strong>of</strong> electrolyte,<br />

oligomeric structures (protein clusters) are formed in<br />

a series <strong>of</strong> thermodynamically unfavorable assembly steps<br />

followed by a growth phase in which clusters are elongated<br />

by further addition <strong>of</strong> protein monomers <strong>and</strong>/or oligomers<br />

upon mutual interaction.<br />

Fitting <strong>of</strong> the time ThT fluorescence evolution shows us<br />

that under acidic conditions, the self-assembly process<br />

becomes more cooperative, as indicated by the values <strong>of</strong><br />

n > 1. This indicates the different nature <strong>of</strong> the self-assembly<br />

pathway <strong>of</strong> HSA under acidic conditions with respect to<br />

physiological pH, since interactions between protein molecules<br />

are modulated by changes in both the pH <strong>and</strong> the initial<br />

protein structure (62). In addition, the plot in the absence <strong>of</strong><br />

electrolyte is fitted in two steps: 1), the fast formation <strong>of</strong><br />

small clusters; <strong>and</strong> 2), the lag phase, which can originate<br />

from the necessity to reach a critical concentration <strong>of</strong> clusters<br />

for aggregation to continue. Probably, oligomeric species<br />

formed in very early stages <strong>of</strong> the aggregation process<br />

(whose existence is indicated by the slight increase <strong>of</strong> ThT<br />

fluorescence at very short incubation times) are more soluble<br />

under acidic conditions, <strong>and</strong> only after they achieve a critical<br />

concentration are they able to grow to generate larger aggregates<br />

(63). However, complete fibril formation can be<br />

achieved only in the presence <strong>of</strong> electrolyte.<br />

Nucleation-dependent growth mechanism<br />

in acidic medium<br />

To corroborate the origin <strong>of</strong> the plateau region in the absence<br />

<strong>of</strong> electrolyte at elevated temperature, we performed a seeding<br />

fibril growth under the conditions previously specified. When<br />

seeds are added to the protein solution, a continuous increase<br />

in ThT fluorescence is observed. This is the typical behavior<br />

observed for a nucleation-type growth mechanism <strong>and</strong> corroborates<br />

the existence <strong>of</strong> this lag phase (see Fig. S6). On the other<br />

h<strong>and</strong>, no changes are observed when protein seeds are added<br />

Biophysical Journal 96(6) 2353–2370<br />

FIGURE 8 Time evolution <strong>of</strong> ThT fluorescence in HSA<br />

solutions incubated at pH 3.0 at 25 C(a <strong>and</strong> b) or65C<br />

(c <strong>and</strong> d) in the absence <strong>and</strong> presence <strong>of</strong> 50 mM NaCl,<br />

respectively.<br />

to a solution containing 50 mM NaCl. The difference between<br />

both solution conditions may arise from the greater hydrophobicity<br />

<strong>of</strong> oligomeric species formed during very earlier incubation<br />

stages in the presence <strong>of</strong> excess electrolyte. In its absence,<br />

electrostatic repulsion between oligomeric species seems to<br />

preclude for some time the formation <strong>of</strong> nuclei with a critical<br />

size to overcome the energy barrier that impedes aggregation.<br />

Structural changes at acidic pH: secondary<br />

structure<br />

At acidic pH <strong>and</strong> room temperature, both minima in [q] at 208<br />

<strong>and</strong> 222 nm are still present before incubation proceeds, which<br />

indicates an important retention <strong>of</strong> this type <strong>of</strong> structure, as<br />

previously described (25,64). During incubation, a slight<br />

decrease in [q] is observed in both the absence <strong>and</strong> presence<br />

<strong>of</strong> added electrolyte, which is compatible with a small<br />

decrease in a-helices <strong>and</strong> the development <strong>of</strong> a small amount<br />

<strong>of</strong> b-sheet conformation due to the appearance <strong>of</strong> small amorphous<br />

aggregates in solution (figure not shown). In contrast,<br />

a shift <strong>of</strong> the 208 nm minimum to lower wavelengths takes<br />

place as incubation proceeds (0–200 h) at 65 C in the absence<br />

<strong>of</strong> electrolyte (Fig. 9 a). An increase in r<strong>and</strong>om coil structure<br />

(characterized by a single minimum below 200 nm) can<br />

account for this shift. Upon further incubation, an additional<br />

red shift occurs as a consequence <strong>of</strong> the increase <strong>of</strong> b-sheet<br />

structure in the aggregates formed, as also indicated by the<br />

increase in ThT fluorescence. The far-UV CD curves at<br />

65 C in the presence <strong>of</strong> NaCl followed a trend similar to those<br />

obtained at physiological pH, although the ellipticity decrease<br />

is less severe at long incubation times as a consequence <strong>of</strong><br />

a lower amount <strong>of</strong> scattered light from fibril aggregates (see<br />

Fig. 9 b). This confirms the lower fibrillar density under these<br />

conditions, which also precludes the formation <strong>of</strong> a fibrillar<br />

gel, in contrast to physiological conditions.<br />

The secondary structure compositions in acidic solution at<br />

room temperature at the beginning <strong>of</strong> the incubation process<br />

156


Fibrillation Pathway <strong>of</strong> HSA 2363<br />

FIGURE 9 Far-UV spectra <strong>of</strong> HSA solutions at 65 C at pH 3.0 in (a) the<br />

absence <strong>of</strong> electrolyte at 1), 0 h; 2), 24 h; 3), 125 h; 4), 175 h; 5), 200 h;<br />

<strong>and</strong> 6), 250 h <strong>of</strong> incubation; <strong>and</strong> (b) the presence <strong>of</strong> 50 mM NaCl at 1),<br />

0 h; 2), 15 h; 3), 48 h; 4), 100 h; 5), 150 h; <strong>and</strong> 6), 200 h <strong>of</strong> incubation.<br />

indicate a reduction in a-helix conformation (~45%) <strong>and</strong> an<br />

increase in b-sheet, b-upturn, <strong>and</strong> coil contents (7%, 20%,<br />

<strong>and</strong> 28%, respectively), typical <strong>of</strong> the acid-exp<strong>and</strong>ed E-state<br />

<strong>of</strong> HSA (25) (Fig. 10). No significant changes were detected<br />

in structural composition during the incubation procedure at<br />

room temperature in either the absence or presence <strong>of</strong> added<br />

electrolyte up to 100 h incubation. At this stage, a slight<br />

increase in b-sheet <strong>and</strong> unordered conformations is observed<br />

for both solution conditions. In contrast, when the temperature<br />

is raised to 65 C, an important increase in b-sheets at<br />

the expense <strong>of</strong> helical conformation occurs as also observed<br />

at physiological pH. This is also accompanied by an increase<br />

in unordered conformation at early incubation times (0–150<br />

h). The change in secondary structure occurs during a longer<br />

time interval (up to 9 days) in acidic solution. The final ahelix<br />

<strong>and</strong> b-sheet compositions are respectively 24% <strong>and</strong><br />

16% in the absence <strong>of</strong> electrolyte (17% <strong>and</strong> 21% in the presence<br />

<strong>of</strong> 50 mM NaCl), compared to 20% <strong>and</strong> 21% at neutral<br />

pH. This indicates small compositional changes in the resulting<br />

amyloid aggregates. The turn conformation also shows<br />

little changes throughout incubation.<br />

FTIR experiments corroborate the structural changes<br />

undergone by the protein molecules as incubation proceeds<br />

at acidic conditions. Before incubation, the amide I b<strong>and</strong><br />

at 1650 cm 1 <strong>and</strong> the amide II b<strong>and</strong> at 1542 cm 1 confirm<br />

that there is still a significant amount <strong>of</strong> a-helices. Incubation<br />

at room temperature under acidic conditions leads to a certain<br />

increase <strong>of</strong> the peak located at ~1627 nm, which corresponds<br />

to intramolecular b-sheet structure, in agreement with<br />

ThT fluorescence <strong>and</strong> CD data. On the other h<strong>and</strong>, after<br />

incubation at 65 C, a red shift <strong>of</strong> the amide I b<strong>and</strong> from<br />

1650 to 1656 cm 1 <strong>and</strong> a blue shift <strong>of</strong> the amide II b<strong>and</strong> to<br />

1540 cm 1 are indicative <strong>of</strong> an increased amount <strong>of</strong> disordered<br />

structure. In addition, the shift <strong>and</strong> further increase <strong>of</strong><br />

the 1627 cm 1 peak to 1624 cm 1 also points to a structural<br />

transformation from an intramolecular hydrogen-bonded<br />

b-sheet to an intermolecular hydrogen-bonded b-sheet structure<br />

(46), as seen for physiological pH. Spectra also show<br />

a small high-frequency component (~1692 cm 1 ) that would<br />

suggest the presence <strong>of</strong> antiparallel b-sheet (47) (see Fig. S7).<br />

Structural changes at acidic pH: tertiary structure<br />

When the pH is decreased, there is an increase in [q]between<br />

260 <strong>and</strong> 280 nm, <strong>and</strong> a slight decrease between 285 <strong>and</strong><br />

300 nm, denoting loss <strong>of</strong> tertiary structure. Nevertheless, there<br />

are still significant CD signals left, suggesting a remaining<br />

tertiary structure, in agreement with previous reports (25,65).<br />

These changes at acidic pH are related to structural rearrangements<br />

<strong>of</strong> all HSA domains. In particular, some increase in ellipticity<br />

below 295 nm takes place during incubation at room<br />

temperature, which points to little further tertiary structural<br />

changes as aggregation takes place; in particular, a loss <strong>of</strong><br />

fine structure is detectable in the 270–295 nm region (see<br />

Fig. S8 a). At 65 C, changes in tertiary structure are more<br />

important; in particular, an almost complete absence <strong>of</strong> the<br />

minima is observed between 260 <strong>and</strong> 270 nm, indicating<br />

a further loss <strong>of</strong> asymmetry around disulfide bridges <strong>and</strong>/or<br />

aromatic residues as incubation proceeds. An additional loss<br />

<strong>of</strong> fine structure in the range <strong>of</strong> 280–295 nm, similar to that<br />

observed at physiological pH, is also detected (Fig. S8, b <strong>and</strong> c).<br />

Upon incubation at acidic pH <strong>and</strong> room temperature,<br />

a certain decrease in the tryptophanyl fluorescence <strong>and</strong> a slight<br />

blue shift <strong>of</strong> the emission maximum occur between days 0 <strong>and</strong><br />

8 <strong>of</strong> incubation, which points to a certain internalization <strong>of</strong><br />

Tryp to the nonpolar environment <strong>of</strong> domain II as a result <strong>of</strong><br />

certain aggregation under these conditions (see Fig. S9).<br />

Because changes in far- <strong>and</strong> near-UV CD spectra are relatively<br />

small for this incubation period, this leads us to think that<br />

structural changes in domain II <strong>of</strong> HSA are involved in this<br />

aggregation process (66). At 65 C, the tryptophanyl residues<br />

are in a more solvent-exposed environment during the incubation<br />

because the fluorescence intensity abruptly decreases<br />

Biophysical Journal 96(6) 2353–2370<br />

157


2364 Juárez et al.<br />

during the first 4 days <strong>of</strong> incubation as a result <strong>of</strong> the sequential<br />

unfolding <strong>of</strong> domains I <strong>and</strong> II <strong>of</strong> HSA. This period <strong>of</strong> time is<br />

slightly shorter than at pH 7.4 (mainly in the presence <strong>of</strong> electrolyte<br />

excess) because the exp<strong>and</strong>ed E state already involves<br />

an important alteration in the tertiary structure.<br />

In situ observation <strong>of</strong> fibrillar structures at acidic<br />

pH: TEM<br />

HSA samples obtained at different incubation times were<br />

also subjected to TEM analysis under acidic conditions. At<br />

room temperature in the absence <strong>of</strong> salt, no aggregates<br />

Biophysical Journal 96(6) 2353–2370<br />

FIGURE 10 Time evolution <strong>of</strong> secondary structure<br />

compositions <strong>of</strong> HSA solutions at pH 3.0 at 25 C(a <strong>and</strong><br />

b) or65C(c <strong>and</strong> d) in the absence <strong>and</strong> the presence <strong>of</strong><br />

50 mM NaCl, respectively. (:) a-helix, (-) b-turn,<br />

(C) unordered, <strong>and</strong> (;) b-sheet conformations.<br />

were observed during the first 2 days <strong>of</strong> the incubation<br />

process. From the third day, the presence <strong>of</strong> small clusters<br />

(globules) <strong>of</strong> aggregated protein could be observed. These<br />

aggregates have a globular or spherical shape, not the regular<br />

fibrillar appearance associated with amyloid structures (see<br />

Fig. 11 a). In addition, less numerous, more elongated aggregates<br />

(20–30 nm long, 3–4 nm wide) can be also observed,<br />

<strong>and</strong> their population slightly increases as incubation<br />

proceeds. The formation <strong>of</strong> these types <strong>of</strong> aggregates is characterized<br />

by a decrease <strong>of</strong> helical structure accompanied by<br />

a slight rise in b-sheet <strong>and</strong> unordered conformations from<br />

CD measurements at long incubation times.<br />

FIGURE 11 TEM pictures <strong>of</strong> the different stages <strong>of</strong> the<br />

HSA fibrillation process at pH 3.0 in the presence <strong>of</strong> 50<br />

mM NaCl (a) at25C after 150 h <strong>of</strong> incubation, <strong>and</strong> at<br />

65 C after (b) 24h,(c) 150 h, <strong>and</strong> (d) 250 h <strong>of</strong> incubation<br />

in the presence <strong>of</strong> 50 mM NaCl.<br />

158


Fibrillation Pathway <strong>of</strong> HSA 2365<br />

When the temperature is raised to 65 C, fibril formation<br />

takes place in several steps. First, formation <strong>of</strong> quasi-spherical<br />

aggregates takes place under incubation times similar to<br />

those at pH 7.4, <strong>and</strong> these seem to be fairly soluble. On the<br />

other h<strong>and</strong>, we have found that a coexistence <strong>of</strong> spherical<br />

aggregates with circular ring-shaped particles <strong>of</strong> larger size<br />

(~100–300 nm) is observed at short incubation times (on<br />

the order <strong>of</strong> a few hours) in the presence <strong>of</strong> excess electrolyte<br />

(Fig. 11 b). These structures appear as a possible intermediate<br />

structural rearrangement <strong>of</strong> smaller protein aggregates<br />

before fibril formation facilitated by electrostatic screening.<br />

There is no evidence as to whether structural reorganization<br />

to form fibrils takes place within this type <strong>of</strong> aggregates, or<br />

whether dissociation <strong>of</strong> HSA molecules from these structures<br />

occurs. Based on our CD data, which indicate that a progressive<br />

gaining <strong>of</strong> b-sheet structure <strong>and</strong> unordered conformation<br />

at the expense <strong>of</strong> a-helices takes place in the first part <strong>of</strong> the<br />

incubation process, we suggest that only after a certain critical<br />

amount <strong>of</strong> b-sheet structure is reached inside these ringshaped<br />

particles will short fibrils be formed in solution from<br />

decomposition <strong>of</strong> the former <strong>and</strong> become stable. Once sufficient<br />

molecules are present within the oligomer, reorganization<br />

steps become thermodynamically favorable as a result <strong>of</strong><br />

an increase in the number <strong>of</strong> hydrogen bonds <strong>and</strong> other stabilizing<br />

interactions. Once fragments <strong>of</strong> highly ordered aggregates<br />

are present, the free energy for addition <strong>of</strong> monomeric<br />

molecules to a growing fibril will become more favorable.<br />

Similar circular ring-shaped structures have also been<br />

observed as a structural intermediate before the formation<br />

<strong>of</strong> amyloid fibrils in the self-assembly process <strong>of</strong> insulin<br />

(67), Ab17-42 peptide (68), Ab1-4 peptide, <strong>and</strong> HaPrP23-<br />

144 prion protein (69). In this case, the circular structures<br />

may incorporate into fibrils but also self-aggregate to form<br />

large, amorphous structures (57). Given the large structural<br />

differences between HSA insulin <strong>and</strong> Ab17-42 peptide, <strong>and</strong><br />

their distinct propensity to form amyloid fibrils, it seems<br />

reasonable to think that these circular ring-shaped structures<br />

can be a sort <strong>of</strong> common structural intermediate <strong>of</strong> amyloid<br />

fibril formation under different solution conditions.<br />

After longer incubation times (6 days) for HSA solutions in<br />

the absence <strong>of</strong> electrolyte at 65 C, small, well-defined, short<br />

prot<strong>of</strong>ibrils (~100 nm long <strong>and</strong> 3–4 nm wide) start to appear<br />

(Fig. 11 c). The corresponding CD spectrum shows a progressive<br />

increase in the proportion <strong>of</strong> b-structure at the expense <strong>of</strong><br />

the helical one during incubation. In contrast, the presence <strong>of</strong><br />

electrolyte involves an additional step: the formation <strong>of</strong> short<br />

curly fibers occurs in a broader incubation timescale than at<br />

physiological pH (Fig. 11 d); under further incubation, the<br />

fibrils appear to increase slightly in number <strong>and</strong> length.<br />

They range from ~100 nm to several hundred nanometers in<br />

length <strong>and</strong> 8 to 11 nm in width; thus, they are shorter on<br />

average than those obtained at physiological conditions but<br />

have the same average width. Closed fibril loops with diameters<br />

<strong>of</strong> ~100 nm were also frequently observed because the<br />

formed filaments can remain short <strong>and</strong> thin to enable them<br />

to bend <strong>and</strong> form closed rings, as also observed for b2 microglobulin<br />

(58), a-crystallin (56), equine lysozyme (70), insulin<br />

(67), <strong>and</strong> a-synuclein (71). This fibrillar material appears<br />

similar to prot<strong>of</strong>ilaments observed in the early stages <strong>of</strong> other<br />

amyloidogenic systems, given the small length <strong>and</strong> the<br />

absence <strong>of</strong> higher-order structures.<br />

DISCUSSION<br />

The ability to form amyloid fibrils is a generic property <strong>of</strong><br />

polypeptide chains, although the propensity <strong>of</strong> different<br />

regions <strong>of</strong> proteins to form such structures varies substantially<br />

(72). The properties <strong>of</strong> unfolded polypeptides, including their<br />

relative propensities for a- <strong>and</strong> b-structure, their intrinsic<br />

solubility, <strong>and</strong> the nature <strong>of</strong> the interactions within the resultant<br />

fibrillar structures, are likely to be particularly important<br />

determinants in their relative abilities to form fibrils. The<br />

conformation <strong>of</strong> the partially folded state is not by itself a critical<br />

feature <strong>of</strong> fibril formation; rather, it is suggested that the<br />

basis for amyloidogenesis is the presence <strong>of</strong> partially denaturation<br />

conditions that destabilize the native fold <strong>of</strong> the protein<br />

but do not preclude noncovalent interactions between the<br />

various groups within the protein. In our case, it appears<br />

necessary to destabilize HSA molecules <strong>and</strong> hence to generate<br />

a partially folded intermediate that can aggregate to form<br />

fibrils. Thus, one can readily atttain conditions in which<br />

such aggregation-prone intermediate states are significantly<br />

populated by lowering the pH or raising the temperature.<br />

ThT fluorescence, CR staining, XRD, <strong>and</strong> TEM pictures<br />

demonstrate the formation <strong>of</strong> amyloid-like structures under<br />

these conditions. The aggregation process is governed by<br />

the balance between attractive <strong>and</strong> repulsive interactions<br />

between protein molecules. Conformational changes induced<br />

by heat increase the number <strong>of</strong> hydrophobic residues exposed<br />

to the aqueous solvent. The exposure <strong>of</strong> these groups results in<br />

attractive hydrophobic interactions that play a dominant role<br />

in the aggregation process. Repulsive forces are induced by<br />

a surface charge that can be modulated by changes in pH,<br />

which controls the net charge <strong>of</strong> the protein, <strong>and</strong> by the ionic<br />

strength <strong>of</strong> the solvent, which controls the screening <strong>of</strong> electrostatic<br />

interactions (73).<br />

Absence <strong>of</strong> a lag phase in HSA fibrillation process<br />

in most solution conditions<br />

A first characteristic <strong>of</strong> the HSA fibrillation process is the<br />

absence <strong>of</strong> a lag phase, as previously observed for bovine<br />

serum albumin (BSA) (74) <strong>and</strong> acyl phosphatase (75) under<br />

all solutions conditions analyzed except for acidic pH at<br />

room temperature. This occurs if the initial aggregation is<br />

a downhill process that does not require a highly organized<br />

<strong>and</strong> unstable nucleus. This is supported by the fact that<br />

seeding <strong>of</strong> preformed aggregates does not accelerate the<br />

fibrillation process. In this regard, it has been suggested<br />

that large multidomain proteins like BSA are able to form<br />

Biophysical Journal 96(6) 2353–2370<br />

159


2366 Juárez et al.<br />

propagation-competent nucleus-like structures (oligomeric<br />

structures) (74). In our case, TEM pictures also show the<br />

formation <strong>of</strong> spherical oligomers upon very short incubation<br />

times, which occur by a means <strong>of</strong> classical coagulation mechanism.<br />

In contrast, the presence <strong>of</strong> a certain lag phase upon<br />

incubation in acidic medium at 65 C in the absence <strong>of</strong> electrolyte<br />

may well indicate that oligomeric aggregates need more<br />

time to develop <strong>and</strong>/or persist for longer times because <strong>of</strong> their<br />

enhanced solubility, so they need to reach a certain number or<br />

size to change the energy l<strong>and</strong>scape <strong>of</strong> the system <strong>and</strong> promote<br />

further aggregation.<br />

Existence <strong>of</strong> different intermediates on the HSA<br />

fibrillation pathway<br />

On a macroscopic scale, the different steps in the fibril formation<br />

pattern, as observed by TEM, consist <strong>of</strong> the formation <strong>of</strong><br />

nonfibrillar aggregates (oligomeric globules) <strong>and</strong> their subsequent<br />

elongation (bead-like structures <strong>and</strong> circular ringshaped<br />

structures), <strong>and</strong> the development <strong>of</strong> prot<strong>of</strong>ilaments<br />

<strong>and</strong> their assembly in fibrils, which can rearrange in more<br />

complex structures. On a molecular level, CD <strong>and</strong> FTIR results<br />

show that HSA possesses native-like a-helical characteristics<br />

with residual b-sheet content before acidic or heat treatment.<br />

In very early periods <strong>of</strong> incubation, HSA forms small,<br />

soluble, globular oligomers <strong>of</strong> mainly native-like molecules<br />

in acidic, physiological, <strong>and</strong>/or high-temperature conditions,<br />

with a progressive increase in b-sheet content <strong>and</strong> unordered<br />

conformation upon further incubation, as revealed by ThT<br />

fluorescence <strong>and</strong> far-UV CD. Thus, upon further incubation<br />

(5–50 h) at elevated temperature, the spectroscopic characteristics<br />

indicate losses <strong>of</strong> persistent tertiary structure along with<br />

unfolding <strong>of</strong> certain secondary structure to different extents<br />

depending on the solution conditions (acidic or physiological<br />

pH, added salt). Different conditions may cause different<br />

regions <strong>of</strong> the polypeptide chains that are relatively flexible<br />

<strong>and</strong> not involved in strong intramolecular interactions<br />

(76,77) to enhance the aggregation process, leading to an<br />

evolution <strong>of</strong> the previous globules into additional intermediate<br />

structures (bead-like structures <strong>and</strong> circular ring-shaped structures<br />

at physiological <strong>and</strong> acidic pH, respectively, with the<br />

latter found only at high ionic strength). Bead-like structures<br />

arise from what appears to be attractive interactions between<br />

globular protein oligomeric clusters, which may result from<br />

an increased exposure <strong>of</strong> hydrophobic residues, mainly in<br />

the presence <strong>of</strong> electrolyte. We found no evidence <strong>of</strong> formation<br />

<strong>of</strong> elongated structures by longitudinal fusion <strong>of</strong> oligomers.<br />

These bead-like structures progressively become more elongated<br />

upon incubation (35 h) due to mutual interactions<br />

between them <strong>and</strong> convert into prot<strong>of</strong>ibrils, in agreement<br />

with a decrease in helical structure as indicated by CD data.<br />

The decrease in diameter accompanying elongation may be<br />

explained on the basis <strong>of</strong> reorganization <strong>of</strong> structure, in particular<br />

<strong>of</strong> the b-str<strong>and</strong>s. Several authors have also reported the<br />

tendency <strong>of</strong> these bead-like structures to transform into fibrillar<br />

Biophysical Journal 96(6) 2353–2370<br />

structures at elevated temperatures caused by partial unfolding<br />

<strong>of</strong> the protein molecules <strong>and</strong> giving rise to conditions conducive<br />

to fibril formation (55–57), as shown in Fig. 6 d.<br />

The combination <strong>of</strong> experimental observations described<br />

here indicates that the formation <strong>of</strong> fibrils from soluble HSA<br />

molecules proceeds in a series <strong>of</strong> stages, the first <strong>of</strong> which is<br />

effectively the presence <strong>of</strong> oligomeric globules. Thus, the<br />

role <strong>of</strong> association-competent oligomeric intermediates may<br />

result in a kinetic pathway via clustering <strong>of</strong> these oligomeric<br />

species, which can be rationalized in the light <strong>of</strong> colloid coagulation<br />

theory, i.e., the formation <strong>of</strong> a critical oligomer or an<br />

ensemble <strong>of</strong> critical oligomers <strong>and</strong> subsequent aggregation<br />

into bead-like structures, <strong>and</strong> then prot<strong>of</strong>ibrillar structures<br />

(78). The persistence <strong>of</strong> these spherical oligomers in solution<br />

coinciding with fibril assembly also supports the view that<br />

they may be ‘‘on-pathway’’ intermediates (see Fig. 6 <strong>and</strong><br />

Fig. S10). Spherical oligomeric structures have been<br />

proposed to serve a key, on-pathway role in both the formation<br />

<strong>and</strong> elongation <strong>of</strong> amyloid fibrils <strong>of</strong> the Sup35 (79) <strong>and</strong><br />

Ure2p yeast prion proteins (80).<br />

In contrast, we speculate that the circular ring-shaped structures<br />

found in acidic medium at high ionic strength, which<br />

appear to be composed <strong>of</strong> two semicircular units, may come<br />

from bending <strong>and</strong> association <strong>of</strong> early-formed, short, beadlike<br />

structures due to a decrease in their solubility. This<br />

decrease would stem from the electrolyte concentration<br />

present in solution, which screens electrostatic interactions<br />

between aggregates. Once sufficient molecules are present<br />

within this type <strong>of</strong> intermediate structure, reorganization steps<br />

become thermodynamically favorable as a result <strong>of</strong> an increase<br />

in the number <strong>of</strong> hydrogen bonds <strong>and</strong> other stabilizing interactions.<br />

Once critical fragments <strong>of</strong> ordered aggregates are<br />

present (with a critical amount <strong>of</strong> b-sheet structure), the free<br />

energy for addition <strong>of</strong> monomeric molecules to a growing<br />

elongated structure becomes more favorable. In this way, short<br />

prot<strong>of</strong>ibrils will be formed in solution upon dissolution <strong>of</strong> the<br />

ring-shaped intermediate structures <strong>and</strong> will become stable.<br />

These intermediates are not observed by TEM once prot<strong>of</strong>ibrils<br />

start to be observed, which suggests that they may act<br />

as reservoirs <strong>of</strong> the initially very short prot<strong>of</strong>ibrils. However,<br />

a deeper structural analysis <strong>of</strong> this structural intermediate<br />

<strong>and</strong> its evolution to prot<strong>of</strong>ibrils is needed, <strong>and</strong> is currently<br />

under way. On the other h<strong>and</strong>, the fact that some elongated<br />

structures (but not fibrils) are formed in acidic pH at room<br />

temperature, in contrast to physiological medium, for which<br />

no aggregation is observed except for some amorphous aggregates,<br />

suggests that the charge distribution on the protein influences<br />

the propensity to form amyloid fibrils. The neutralization<br />

<strong>of</strong> negative charges at acidic pH may promote this elongation.<br />

HSA fibrils are curly<br />

The thermodynamic instability <strong>of</strong> prefibrillar aggregates<br />

causes them to evolve into more stable assemblies upon additional<br />

incubation (>75 h), eventually leading to the appearance<br />

160


Fibrillation Pathway <strong>of</strong> HSA 2367<br />

<strong>of</strong> stable fibrils, mainly at physiological conditions at elevated<br />

temperatures, where structural reorganization results in the<br />

large majority <strong>of</strong> hydrophobic groups being concealed from<br />

the solvent. This is corroborated by ThT fluorescence, far<br />

UV-CD, <strong>and</strong> FTIR spectra, which demonstrate that the major<br />

elements <strong>of</strong> ordered secondary structure are b-sheets, suggesting<br />

that the a-helical regions <strong>of</strong> the native protein have undergone<br />

significant structural changes. TEM pictures corroborate<br />

this picture, showing that the fibrils formed are mainly, narrow,<br />

branched, <strong>and</strong> elongated but curly, in contrast to the straight<br />

needle-like structures characteristic <strong>of</strong> bona fide fibrils. As<br />

incubation proceeds, these worm-like structures increase in<br />

length. These fibrils display the typical 4.8 A˚ peak indicating<br />

the typical interstr<strong>and</strong> distance <strong>of</strong> classical fibrils, <strong>and</strong> the<br />

11 A˚ equatorial reflection corresponds to the intersheet<br />

spacing, with the b-sheets stacked face to face to form the<br />

core structure <strong>of</strong> prot<strong>of</strong>ilaments (81). Based on the CD <strong>and</strong><br />

FTIR data, we speculate that these types <strong>of</strong> fibrils may be<br />

formed by a seam <strong>of</strong> b-sheet structure decorated by relatively<br />

disorganized a-helical structure, as previously observed for<br />

RNase A (82) or yeast Ure2p (83). In addition, an antiparallel<br />

arrangement <strong>of</strong> b-str<strong>and</strong>s forming the b-sheet structure <strong>of</strong> the<br />

HSA fibrils seems to be the most probable configuration, as<br />

denoted by the b<strong>and</strong> at ~1691–1693 cm 1 in the FTIR spectra.<br />

Curly aggregates have been also seen in other proteins, such as<br />

b2 microglobulin (58) <strong>and</strong> a-crystallin (56). In addition, intramolecular<br />

end-to-end association <strong>of</strong> short individual filaments<br />

appears to be favorable, as TEM images reveal the presence <strong>of</strong><br />

a substantial number <strong>of</strong> closed loops appearing to form spontaneously,<br />

mainly in the absence <strong>of</strong> electrolyte at acidic conditions.<br />

Qualitatively, the probability <strong>of</strong> a single fibril joining<br />

end-to-end to form a closed loop will be high if the fibril is<br />

sufficiently flexible <strong>and</strong> <strong>of</strong> appropriate length for the ends to<br />

find themselves regularly in close proximity to each other<br />

(84). Indeed, the results presented here exemplify the favorable<br />

nature <strong>of</strong> loop formation when such fibril morphologies are<br />

adopted. Loop formation was reported previously for other<br />

amyloid-forming systems (85,86).<br />

On the other h<strong>and</strong>, it is interesting to note that a similar<br />

curly morphology for HSA can be achieved under different<br />

solution conditions: the single filament, in the form <strong>of</strong> both<br />

open flexible chains <strong>and</strong> closed loops, is observed at both<br />

physiological <strong>and</strong> acidic pH in the presence <strong>of</strong> electrolyte.<br />

Nevertheless, the rate <strong>and</strong> extent <strong>of</strong> aggregation depends<br />

on the solution conditions: the amount <strong>of</strong> formed fibrils<br />

<strong>and</strong> their length (larger at physiological conditions) favor<br />

interactions between early aggregates to form fibrils. This<br />

is revealed by the formation <strong>of</strong> a gel phase under suitable<br />

solution conditions.<br />

Curly fibers can evolve into a suprafibrillar<br />

structure<br />

Very long incubation times (>150 h) lead to a more complex<br />

morphological variability among amyloid fibrils (e.g., long<br />

straight fibrils, flat-ribbon structures, or laterally connected<br />

fibers). These compact, mature fibrillar assemblies formed<br />

at the endpoint <strong>of</strong> the aggregation process may result from<br />

an effort to minimize the exposure <strong>of</strong> hydrophobic residues,<br />

<strong>and</strong> are also likely to result in increased van der Waals interactions,<br />

leading to greater stability, as previously noted for<br />

b2-microglobulin (87), a-synuclein (5), <strong>and</strong> insulin (88). A<br />

similar progression in structure from curly fibrils to mature<br />

straight fibers was also observed for Ab-peptide (89) <strong>and</strong><br />

insulin (90). In contrast, BSA fibrillation was shown to be<br />

halted at the early curly stage, despite the enormous structural<br />

similarities with HSA, since no further development<br />

in fibrillar structure over long timescales was observed<br />

(74). Mature fibrils are thicker <strong>and</strong> stiffer than single fibrils<br />

<strong>and</strong> appear to be formed by lateral (side-by-side) assembly<br />

<strong>of</strong> two or more individual filaments. Nevertheless, under<br />

physiological conditions, we observed that mature straight<br />

fibrils can also grow by swollen protein particles tending<br />

to align within the immediate vicinity <strong>of</strong> the fibers, as shown<br />

in Fig. 6 i, serving the single fiber as a lateral template or<br />

scaffold for small protein molecules, <strong>and</strong> would constitute<br />

a subcomponent in mature fibrils. Their population is dependent<br />

on solution conditions <strong>and</strong> lateral interactions between<br />

fibrils. This may be an additional pathway to the formation<br />

<strong>of</strong> mature fibrils via association <strong>of</strong> prot<strong>of</strong>ibrils. Conformational<br />

preferences for a certain pathway to become active<br />

may exist, <strong>and</strong> thus the influence <strong>of</strong> environmental conditions<br />

such as pH, temperature, <strong>and</strong> salt must be considered.<br />

Thus, it seems that a single filament may act as a ‘‘lateral<br />

template’’ or scaffold for small protein particles, which<br />

would constitute a neighboring subcord-like feature in the<br />

fiber shown in Fig. 6 h. This filament þ monomers/oligomers<br />

scenario is an alternative pathway to the otherwise<br />

dominating filament þ filament manner <strong>of</strong> the protein fibril’s<br />

lateral growth, as has been also observed for insulin (57).<br />

AFM data recently reported by Green et al. (91) suggest<br />

that mature fibrils from human amylin are unlikely to be<br />

assembled by the lateral association <strong>of</strong> prot<strong>of</strong>ibrils.<br />

It appears, therefore, that even for a homogeneous HSA<br />

sample undergoing uniform temperature or acidic treatment,<br />

there is still more than one mode <strong>of</strong> assembly <strong>of</strong> filaments<br />

(92). Such polymorphism may be caused by differences in<br />

the number <strong>of</strong> filaments assembled in the mature fibrils;<br />

however, it may also result from the incorporation in different<br />

regions <strong>of</strong> the sequence <strong>of</strong> the polypeptide chain with various<br />

types <strong>of</strong> fibrils. In addition, the differences between the populations<br />

<strong>of</strong> fibers clearly involve not only the rate <strong>of</strong> the aggregation<br />

process, but also the different quaternary folds.<br />

Although the different distribution pr<strong>of</strong>iles <strong>of</strong> the fibrillar<br />

features may still be explained in terms <strong>of</strong> kinetic effects<br />

(e.g., temperature may differently affect the kinetics at various<br />

stages <strong>of</strong> the assembly, effectively marginalizing certain<br />

sequential processes), high temperatures may also have<br />

a more direct effect on amyloidogenesis, for instance, by<br />

increasing the thermal energy <strong>of</strong> the interacting molecules<br />

Biophysical Journal 96(6) 2353–2370<br />

161


2368 Juárez et al.<br />

<strong>and</strong> hence causing the alignment <strong>of</strong> filaments to become less<br />

accurate. Such unspecific effects may contribute to morphological<br />

differences in any protein’s amyloid samples induced<br />

at high temperature. In this regard, recent reports based on<br />

nuclear magnetic resonance showed that different fibril<br />

morphologies have different underlying secondary structures,<br />

<strong>and</strong> as such are likely produced by distinct independent<br />

assembly pathways (93,94).<br />

A summary <strong>of</strong> the various fibril morphologies observed in<br />

this study is shown in Fig. 12 together with a schematic<br />

representation <strong>of</strong> the assembly process. We propose that at<br />

elevated temperature (except at pH 3.0 in the absence <strong>of</strong> electrolyte),<br />

HSA forms rapidly globular oligomers that upon<br />

mutual interaction evolve into more elongated structures<br />

(bead-like) that grow to prot<strong>of</strong>ibrils either by subsequent<br />

annealing <strong>of</strong> oligomers <strong>and</strong>/or protein monomers. Mature<br />

fibrils can be formed by lateral association <strong>of</strong> prot<strong>of</strong>ibrils<br />

or the addition <strong>of</strong> protein oligomers to the growing fibril,<br />

both at the ends <strong>of</strong> the fibril <strong>and</strong> by lateral fusion. Ringlike<br />

structures are present in acidic conditions at elevated<br />

temperature in the presence <strong>of</strong> electrolyte as an additional<br />

intermediate state formed by association <strong>of</strong> short bead-like<br />

structures, which disappears when prot<strong>of</strong>ibrils are observed<br />

in solution. Thus, we think that they may act as reservoirs<br />

<strong>of</strong> initially very short prot<strong>of</strong>ibrils.<br />

CONCLUSIONS<br />

We observed the formation <strong>of</strong> prot<strong>of</strong>ibrils, curly fibers, <strong>and</strong><br />

mature fibrils by the protein HSA under different solution<br />

conditions. We analyzed the fibrillation process <strong>and</strong> the<br />

conformational changes associated with it by using different<br />

spectroscopic techniques, <strong>and</strong> confirmed the necessary development<br />

<strong>of</strong> b-sheet structure upon fibrillation. In addition, the<br />

shapes <strong>of</strong> the different structural intermediates <strong>and</strong> final products<br />

in the fibrillation process were observed by TEM, SEM,<br />

<strong>and</strong> AFM. The obtained fibrils show structural features typical<br />

<strong>of</strong> classical amyloid fibers, as denoted by XRD, CD, <strong>and</strong> fluorescence<br />

spectroscopies, <strong>and</strong> TEM. A model <strong>of</strong> fibril formation<br />

based on the elongation <strong>of</strong> protein oligomers through<br />

mutual interactions <strong>and</strong> subsequent annealing <strong>and</strong> growth is<br />

FIGURE 12 Mechanisms <strong>of</strong> fibril formation for HSA.<br />

Biophysical Journal 96(6) 2353–2370<br />

presented. Nevertheless, some differences in the fibrillation<br />

mechanism occur depending on the solution conditions; for<br />

example, ring-shaped structures are observed only as a structural<br />

intermediate under acidic conditions in the presence <strong>of</strong><br />

added electrolyte.<br />

SUPPORTING MATERIAL<br />

Ten figures <strong>and</strong> a table are available at http://www.biophysj.org/biophysj/<br />

supplemental/S0006-3495(09)00322-1.<br />

We thank Dr. Eugenio Vázquez for his assistance with the CD measurements.<br />

This study was supported by the Ministerio de Educación y Ciencia (project<br />

MAT-2007-61604).<br />

REFERENCES<br />

1. Reches, M., <strong>and</strong> E. Gazit. 2003. Casting metal nanowires within discrete<br />

self-assembled peptide nanotubes. Science. 300:625–627.<br />

2. Rajagopal, K., <strong>and</strong> J. P. Schneider. 2004. <strong>Self</strong>-assembling peptides <strong>and</strong><br />

proteins for nanotechnological applications. Curr. Opin. Struct. Biol.<br />

14:480–486.<br />

3. Stefani, M., <strong>and</strong> C. M. Dobson. 2003. Protein aggregation <strong>and</strong> aggregate<br />

toxicity: new insights into protein folding, misfolding diseases<br />

<strong>and</strong> biological evolution. J. Mol. Med. 81:678–699.<br />

4. Soto, C., L. Estrada, <strong>and</strong> J. Castilla. 2006. Amyloids, prions <strong>and</strong> the<br />

inherent infectious nature <strong>of</strong> misfolded protein aggregates. Trends<br />

Biochem. Sci. 31:150–155.<br />

5. Khurana, R., C. Ionescu-Zanetti, M. Pope, J. Li, L. Nelson, et al. 2003.<br />

A general mode for amyloid fibril assembly based on morphological<br />

studies using atomic force microscopy. Biophys. J. 85:1135–1144.<br />

6. Chiti, F., M. Stefani, N. Taddei, G. Ramponi, <strong>and</strong> C. M. Dobson. 2003.<br />

Rationalization <strong>of</strong> the effects <strong>of</strong> mutations on peptide <strong>and</strong> protein aggregation<br />

rates. Nature. 424:805–808.<br />

7. Williams, A. D., E. Portelius, I. Kheterpal, J. T. Guo, K. D. Cook, et al.<br />

2004. Mapping a b amyloid fibril secondary structure using scanning<br />

proline mutagenesis. J. Mol. Biol. 335:833–842.<br />

8. Hortscansky, P., T. Christopeit, V. Schroeckh, <strong>and</strong> M. Fändrich. 2005.<br />

Thermodynamic analysis <strong>of</strong> the aggregation propensity <strong>of</strong> oxidized<br />

Alzheimer’s b-amyloid variants. Protein Sci. 14:2915–2918.<br />

9. Fändrich, M., V. Forge, K. Buder, M. Kittler, C. M. Dobson, et al. 2003.<br />

Myoglobin forms amyloid fibrils by association <strong>of</strong> unfolded peptide<br />

segments. Proc. Natl. Acad. Sci. USA. 100:15463–15468.<br />

10. Gazit, E. 2002. The ‘‘correctly folded’’ state <strong>of</strong> proteins: is it a metastable<br />

state? Angew. Chem. Int. Ed. Engl. 41:257–259.<br />

11. Gillmore, J. D., A. J. Stangou, G. A. Tennet, D. R. Booth, J. O’Grady,<br />

et al. 2001. Clinical <strong>and</strong> biochemical outcome <strong>of</strong> hepatorenal transplantation<br />

for hereditary systemic amyloidosis associated with apolipoprotein<br />

AI Gly26Arg. Transplantation. 71:986–992.<br />

12. Carulla, N., G. L. Caddy, D. R. May, J. Zurdo, M. Gairi, et al. 2005.<br />

Molecular recycling within amyloid fibrils. Nature. 436:554–558.<br />

13. Pepys, M. B., J. Herbert, W. L. Hutchison, G. A. Tennent, H. J. Lachmann,<br />

et al. 2002. Targeted pharmacological depletion <strong>of</strong> serum amyloid<br />

P component for treatment <strong>of</strong> human amyloidosis. Nature. 417:254–259.<br />

14. Janus, C., M. A. Chishti, <strong>and</strong> D. Westaway. 2000. Transgenic mouse<br />

models <strong>of</strong> Alzheimer’s disease. Biochim. Biophys. Acta. 1502:63–75.<br />

15. Cohen, F. E., <strong>and</strong> J. W. Kelly. 2003. Therapeutic approaches to proteinmisfolding<br />

diseases. Nature. 426:905–909.<br />

16. Tanaka, M., P. Chien, N. Narber, R. Cooke, <strong>and</strong> J. S. Weissman. 2004.<br />

Conformational variations in an infectious protein determine prion<br />

strain differences. Nature. 428:323–328.<br />

162


Fibrillation Pathway <strong>of</strong> HSA 2369<br />

17. Fändrich, M., <strong>and</strong> C. M. Dobson. 2002. The behavior <strong>of</strong> polyamino<br />

acids reveals an inverse side chain effect in amyloid structure formation.<br />

EMBO J. 21:5682–5690.<br />

18. Gorinstein, S., A. Caspi, A. Rosen, I. Goshev, M. Zemser, et al. 2002.<br />

Structure characterization <strong>of</strong> human serum aproteins in solution <strong>and</strong> dry<br />

state. J. Pept. Res. 59:71–78.<br />

19. Taboada, P., S. Barbosa, E. Castro, <strong>and</strong> V. Mosquera. 2006. Amyloid<br />

fibril formation <strong>and</strong> other aggregate species formed by human serum<br />

albumin association. J. Phys. Chem. B. 110:20733–20736.<br />

20. Pace, C. N., F. Vajdos, L. Fee, G. Grimsley, <strong>and</strong> T. Gray. 1995. How to<br />

measure <strong>and</strong> predict the molar absorption coefficient <strong>of</strong> a protein.<br />

Protein Sci. 4:2411–2423.<br />

21. Sreerama, N., <strong>and</strong> R. W. Woody. 2000. Estimation <strong>of</strong> protein secondary<br />

structure from circular dichroism spectra: comparison <strong>of</strong> CONTIN,<br />

SELCON, <strong>and</strong> CDSSTR methods with an exp<strong>and</strong>ed reference set.<br />

Anal. Biochem. 287:252–260.<br />

22. Uversky, V. N., <strong>and</strong> A. L. Fink. 2004. Conformational constraints for<br />

amyloid fibrillation: the importance <strong>of</strong> being unfolded. Biochim.<br />

Biophys. Acta. 1698:131–153.<br />

23. Calamai, M., F. Chiti, <strong>and</strong> C. M. Dobson. 2005. Amyloid fibril formation<br />

can proceed from different conformations <strong>of</strong> a partially unfolded<br />

protein. Biophys. J. 89:4201–4210.<br />

24. Morel, B., S. Casares, <strong>and</strong> F. A. Conejo-Lara. 2006. Single mutation<br />

induces amyloid aggregation in the a-spectrin SH3 domain: analysis<br />

<strong>of</strong> the early stages <strong>of</strong> fibril formation. J. Mol. Biol. 356:453–468.<br />

25. Dockal, M., D. C. Carter, <strong>and</strong> F. Rüker. 2000. Conformational transitions<br />

<strong>of</strong> the three recombinant domains <strong>of</strong> human serum albumin<br />

depending on pH. J. Biol. Chem. 275:3042–3050.<br />

26. Peters, T. Jr., 1996. All About Albumin: Biochemistry, Genetics <strong>and</strong><br />

Medical Application. Academic Press, New York.<br />

27. Geisow, M. J., <strong>and</strong> G. H. Beaven. 1977. Physical <strong>and</strong> binding properties<br />

<strong>of</strong> large fragments <strong>of</strong> human serum albumin. Biochem. J. 165:477–484.<br />

28. Khan, M. Y. 1986. Direct evidence for the involvement <strong>of</strong> domain III in<br />

the N-F transition <strong>of</strong> bovine serum albumin. Biochem. J. 236:307–310.<br />

29. Ahmad, B., M. K. A. Khan, S. Haq, <strong>and</strong> R. H. Khan. 2004. Intermediate<br />

formation at lower urea concentration in ‘B’ isomer <strong>of</strong> human serum<br />

albumin: a case study using domain specific lig<strong>and</strong>s. Biochem. Biophys.<br />

Res. Commun. 314:166–173.<br />

30. Taboada, P., S. Barbosa, E. Castro, M. Gutiérrez-Pichel, <strong>and</strong> V. Mosquera.<br />

2007. Effect <strong>of</strong> solvation on the structure conformation <strong>of</strong> human<br />

serum albumin in aqueous-alcohol mixed solvents. Chem. Phys.<br />

340:59–68.<br />

31. Farruggia, B., F. Rodriguez, R. Rigatuso, G. Fidelio, <strong>and</strong> G. Picó. 2001.<br />

The participation <strong>of</strong> human serum albumin domains in chemical <strong>and</strong><br />

thermal unfolding. J. Protein Chem. 20:81–89.<br />

32. Flora, K., J. D. Brennan, G. A. Baker, M. A. Doody, <strong>and</strong> F. V. Bright.<br />

1998. Unfolding <strong>of</strong> acrylodan-labeled human serum albumin probed<br />

by steady-state <strong>and</strong> time resolved fluorescence methods. Biophys. J.<br />

75:1084–1096.<br />

33. Rader, A. J., B. M. Hespenheide, K. A. Kuhn, <strong>and</strong> M. F. Thorpe. 2002.<br />

Protein unfolding: rigidity lost. Proc. Natl. Acad. Sci. USA. 99:<br />

3540–3545.<br />

34. LeVine, H. 3rd. , 1993. Thi<strong>of</strong>lavine T interaction with synthetic<br />

Alzheimer’s disease b-amyloid peptides: detection <strong>of</strong> amyloid aggregation<br />

in solution. Protein Sci. 2:404–410.<br />

35. Klunk, W. E., J. W. Pettegrew, <strong>and</strong> D. J. Abraham. 1989. Quantitative<br />

evaluation <strong>of</strong> Congo Red binding to amyloid-like proteins with<br />

a b-pleated sheet conformation. J. Histochem. Cytochem. 37:1273–1281.<br />

36. Khurana, R., C. Coleman, C. Ionescu-Zanetti, S. A. Carter, V. Krishna,<br />

et al. 2005. Mechanisms <strong>of</strong> thi<strong>of</strong>lavin T binding to amyloid fibrils.<br />

J. Struct. Biol. 151:229–238.<br />

37. Sagis, L. M. C., C. Veerman, <strong>and</strong> E. Van der Linden. 2004. Mesoscopic<br />

properties <strong>of</strong> semiflexible amyloid fibrils. Langmuir. 20:924–927.<br />

38. Morozova-Roche, L. A., J. A. Jones, W. Noppe, <strong>and</strong> C. M. Dobson.<br />

1999. Independent nucleation <strong>and</strong> heterogeneous assembly <strong>of</strong> structure<br />

during unfolding <strong>of</strong> equine lysozime. J. Mol. Biol. 289:1055–1073.<br />

39. Jund, P., R. Jullien, <strong>and</strong> I. Campbell. 2001. R<strong>and</strong>om walks on fractals<br />

<strong>and</strong> stretched exponential relaxation. Phys. Rev. E. 63:361311–361314.<br />

40. Harper, J. D., <strong>and</strong> P. T. J. L<strong>and</strong>sbury. 1997. Models <strong>of</strong> amyloid seeding<br />

in Alzheimer’s disease <strong>and</strong> scrapie: mechanistic truths <strong>and</strong> physiological<br />

consequences <strong>of</strong> the time-dependent solubility <strong>of</strong> amyloid proteins.<br />

Annu. Rev. Biochem. 66:385–407.<br />

41. Lomakin, A., D. S. Chung, G. B. Benedek, D. A. Kirschner, <strong>and</strong><br />

D. B. Teplow. 1996. On the nucleation <strong>and</strong> growth <strong>of</strong> amyloid bprotein<br />

fibrils: detection <strong>of</strong> nuclei <strong>and</strong> quantitation <strong>of</strong> rate constants.<br />

Proc. Natl. Acad. Sci. USA. 93:1125–1129.<br />

42. Lomakin, A., D. B. Teplow, D. A. Kirschner, <strong>and</strong> G. B. Benedek. 1997.<br />

Kinetic theory <strong>of</strong> fibrillogenesis <strong>of</strong> amyloid b-protein. Proc. Natl. Acad.<br />

Sci. USA. 94:7942–7947.<br />

43. Ferrone, F. 1999. Analysis <strong>of</strong> protein aggregation kinetics. Methods<br />

Enzymol. 309:256–274.<br />

44. Susi, H., <strong>and</strong> D. M. Byler. 1983. Protein structure by Fourier transform<br />

infrared spectroscopy: second derivative spectra. Biochem. Biophys.<br />

Res. Commun. 115:391–397.<br />

45. Lin, S. -Y., Y. -S. Wei, M. -J. Li, <strong>and</strong> S. -L. Wang. 2004. Effect <strong>of</strong> ethanol<br />

or/<strong>and</strong> captopril on the secondary structure <strong>of</strong> human serum albumin<br />

before <strong>and</strong> after protein binding. Eur. J. Pharm. Biopharm. 57:457–464.<br />

46. Jackson, M., <strong>and</strong> H. H. Mantsch. 1995. The use <strong>and</strong> misuse <strong>of</strong><br />

FTIR spectroscopy in the determination <strong>of</strong> protein structure. Crit.<br />

Rev. Biochem. Mol. Biol. 30:95–120.<br />

47. Casal, H. L., U. Kohler, <strong>and</strong> H. H. Mantsch. 1988. Structural <strong>and</strong><br />

conformational changes <strong>of</strong> b-lactoglobulin B: an infrared spectroscopy<br />

study <strong>of</strong> the effect <strong>of</strong> pH <strong>and</strong> temperature. Biochim. Biophys. Acta.<br />

957:11–20.<br />

48. Fabian, H., L. -P. Choo, G. I. Szendrei, M. Jackson, W. C. Halliday,<br />

et al. 1993. Infrared spectroscopic characterization <strong>of</strong> Alzheimer<br />

plaques. Appl. Spectrosc. 47:1513–1518.<br />

49. Lin, S. -Y., Y. -S. Wei, M. -J. Li, <strong>and</strong> S. -L. Wang. 2004. Effect <strong>of</strong><br />

ethanol or/<strong>and</strong> captopril on the secondary structure <strong>of</strong> human serum<br />

albumin before <strong>and</strong> after protein binding. Spectrochim. Acta [A].<br />

60:3107–3111.<br />

50. Uversky, V. N., N. V. Narizhneva, T. V. Ivanova, <strong>and</strong> A. Y. Tomashevski.<br />

1997. Rigidity <strong>of</strong> human a-fetoprotein tertiary structure is under<br />

lig<strong>and</strong> control. Biochemistry. 494:13638–13645.<br />

51. Muzammil, S., Y. Kumar, <strong>and</strong> S. Tayyab. 1999. Molten globule-like<br />

state <strong>of</strong> human serum albumin at low pH. FEBS Lett. 266:26–32.<br />

52. Qiu, W., L. Zhang, O. Okobiah, Y. Yang, L. Wang, et al. 2006. Ultrafast<br />

solvation dynamics <strong>of</strong> human serum albumin: correlations with conformational<br />

transitions <strong>and</strong> site-selected recognition. J. Phys. Chem. B.<br />

110:10540–10549.<br />

53. Cowgill, R. W. 1968. Fluorescence <strong>and</strong> protein structure. XIV. Tyrosine<br />

fluorescence in helical muscle proteins. Biochim. Biophys. Acta.<br />

161:431–438.<br />

54. Shahi, P., R. Sharma, S. Sanger, I. Kumar, <strong>and</strong> R. S. Jolly. 2007. Formation<br />

<strong>of</strong> amyloid fibrils via longitudinal growth <strong>of</strong> oligomers. Biochemistry.<br />

46:7365–7373.<br />

55. Burgio, M. R., P. M. Bwennet, <strong>and</strong> J. F. Koretz. 2001. Heat-induced<br />

quaternary transitions in hetero- <strong>and</strong> homo-polymers <strong>of</strong> a-crystallin.<br />

Mol. Vis. 7:228–233.<br />

56. Meehan, S., Y. Berry, B. Luisa, C. M. Dobson, J. A. Carver, et al. 2004.<br />

Amyloid fibril formation by lens crystallin proteins <strong>and</strong> its implications<br />

for cataract formation. J. Biol. Chem. 279:3413–3419.<br />

57. Jansen, R., W. Dzwolak, <strong>and</strong> R. Winter. 2005. Amyloidogenic selfassembly<br />

<strong>of</strong> insulin aggregates probed by high resolution atomic force<br />

microscopy. Biophys. J. 88:1344–1353.<br />

58. Radford, S. E., W. S. Gosal, <strong>and</strong> G. W. Platt. 2005. Towards an underst<strong>and</strong>ing<br />

<strong>of</strong> the structural molecular mechanism <strong>of</strong> b2-microglobulin<br />

amyloid formation in vitro. Biochim. Biophys. Acta. 1753:51–63.<br />

59. Kodali, R., <strong>and</strong> R. Wetzel. 2007. Polymosphism in the intermediates<br />

<strong>and</strong> products <strong>of</strong> amyloid assembly. Curr. Opin. Struct. Biol. 17:48–57.<br />

60. Sunde, M., <strong>and</strong> C. Blake. 1997. The structure <strong>of</strong> amyloid fibrils by electron<br />

microscopy <strong>and</strong> X-ray diffraction. Adv. Protein Chem. 50:123–159.<br />

Biophysical Journal 96(6) 2353–2370<br />

163


2370 Juárez et al.<br />

61. Bouchard, M., J. Zurdo, E. J. Nettleton, C. M. Dobson, <strong>and</strong> C. V. Robinson.<br />

2000. Formation <strong>of</strong> insulin amyloid fibrils followed by FTIR<br />

simultaneously with CD <strong>and</strong> electron microscopy. Protein Sci.<br />

9:1960–1967.<br />

62. Vetri, V., F. Librizzi, M. Leone, <strong>and</strong> V. Militello. 2007. Thermal aggregation<br />

<strong>of</strong> bovine serum albumin at different pH: comparison with<br />

human serum albumin. Eur. Biophys. J. 36:717–725.<br />

63. Bader, R., R. Bamford, J. Zurdo, B. F. Luisi, <strong>and</strong> C. M. Dobson. 2006.<br />

Probing the mechanism <strong>of</strong> amyloidogenesis through a t<strong>and</strong>em repeat <strong>of</strong><br />

the PI3–SH3 domain suggests a generis model for protein aggregation<br />

<strong>and</strong> fibril formation. J. Mol. Biol. 356:189–208.<br />

64. Shaw, A. K., <strong>and</strong> S. K. Pal. 2008. Spectroscopic studies on the effect<br />

<strong>of</strong> temperature on pH-induced folded states <strong>of</strong> human serum albumin.<br />

J. Photochem. Photobiol. B. 90:69–77.<br />

65. Ahmad, B., S. Parveen, <strong>and</strong> R. H. Khan. 2006. Effect <strong>of</strong> albumin<br />

conformation on the binding <strong>of</strong> cipr<strong>of</strong>loxacin to human serum albumin:<br />

a novel approach directly assigning binding site. Biomacromolecules.<br />

7:1350–1356.<br />

66. Eftink, M. R., <strong>and</strong> C. A. Ghiron. 1976. Exposure <strong>of</strong> tryptophyl residues<br />

in proteins. Quantitative determination by fluorescence quenching<br />

studies. Biochemistry. 15:672–679.<br />

67. Sibley, S. P., K. Sosinsky, L. E. Gulian, E. J. Gibbs, <strong>and</strong> R. F. Pasternack.<br />

2008. Probing the mechanism <strong>of</strong> insulin aggregation with added<br />

metalloporphyrins. Biochemistry. 47:2858–2865.<br />

68. Zheng, J., H. Jang, B. Ma, <strong>and</strong> R. Nussinov. 2008. Annular structures as<br />

intermediates in fibril formation <strong>of</strong> Alzheimer Ab17–42. J. Phys. Chem.<br />

B. 112:6856–6865.<br />

69. Moore, R. E., S. F. Hayes, E. R. Fischer, <strong>and</strong> S. A. Priola. 2007.<br />

Amyloid formation via supramolecular peptide assemblies. Biochemistry.<br />

46:7079–7087.<br />

70. Malisauskas, M., V. Zamotin, J. Jass, W. Noppe, C. M. Dobson, et al.<br />

2003. Amyloid prot<strong>of</strong>ilaments from the calcium-binding protein equine<br />

lysozyme: formation <strong>of</strong> ring <strong>and</strong> linear structures depends on pH <strong>and</strong><br />

metal ion concentration. J. Mol. Biol. 330:879–890.<br />

71. Lashuel, H. A., D. Hartley, B. M. Petre, T. Walz, <strong>and</strong> P. T. Lansbury Jr.<br />

2002. Neurodegenerative disease: amyloid pores from pathogenic mutations.<br />

Nature. 418:291.<br />

72. Jarrett, J. T., <strong>and</strong> P. T. Lansbury Jr. 1992. Amyloid fibril formation<br />

requires a chemically discriminating nucleation event—studies <strong>of</strong> an<br />

amyloidogenic sequence from the bacterial protein OSMB. Biochemistry.<br />

31:12345–12352.<br />

73. Lefebvre, J., D. Renard, <strong>and</strong> A. C. Sánchez-Gimeno. 1998. Structure<br />

<strong>and</strong> rheology <strong>of</strong> heat-set gels <strong>of</strong> globular proteins. I. Bovine serum<br />

albumin gels in isoelastic conditions. Rheol. Acta. 37:345–357.<br />

74. Holm, N. K., S. K. Jespersen, L. V. Thomassen, T. Y. Wolff, P. Sehgal,<br />

et al. 2007. Aggregation <strong>and</strong> fibrillation <strong>of</strong> bovine serum albumin. Biochim.<br />

Biophys. Acta. 1774:1128–1138.<br />

75. Chiti, F., P. Webster, N. Taddei, A. Clark, M. Stefani, et al. 1999.<br />

Designing conditions for in vitro, formation <strong>of</strong> amyloid prot<strong>of</strong>ilaments<br />

<strong>and</strong> fibrils. Proc. Natl. Acad. Sci. USA. 96:3590–3594.<br />

76. Krishnan, R., <strong>and</strong> S. L. Lindquist. 2005. Structural insights into a yeast<br />

prion illuminate nucleation <strong>and</strong> strain diversity. Nature. 435:765–772.<br />

77. Moraitakis, G., <strong>and</strong> J. M. Goodfellow. 2003. Simulations <strong>of</strong> human lysozime:<br />

probing the conformations triggering amyloidosis. Biophys. J.<br />

84:2149–2158.<br />

Biophysical Journal 96(6) 2353–2370<br />

78. Modler, A. J., K. Gast, G. Lutsch, <strong>and</strong> G. Damaschun. 2003. <strong>Assembly</strong><br />

<strong>of</strong> amyloid prot<strong>of</strong>ibrils via critical oligomers. A novel pathway <strong>of</strong><br />

amyloid formation. J. Mol. Biol. 325:135–148.<br />

79. Serio, T. R., A. G. Cashikar, A. S. Kowal, G. J. Sawicki, J. J. Moleshi,<br />

et al. 2000. Nucleated conformational conversion <strong>and</strong> the replication <strong>of</strong><br />

conformational information by a prion determinant. Science. 289:<br />

1317–1321.<br />

80. Jiang, Y., H. Li, L. Zhu, J. M. Zhou, <strong>and</strong> S. Perrett. 2004. Amyloid<br />

nucleation <strong>and</strong> hierarchical assembly <strong>of</strong> Ure2p fibrils. Role <strong>of</strong> asparagines/glutamine<br />

repeat <strong>and</strong> nonrepeat regions <strong>of</strong> the prion domains.<br />

J. Biol. Chem. 279:3361–3369.<br />

81. Makin, O. S., <strong>and</strong> L. C. Serpell. 2005. X-ray diffraction studies <strong>of</strong><br />

amyloid structure. Methods Mol. Biol. 299:67–80.<br />

82. Sambashivan, S., M. R. Liu, M. R. Sawaya, M. Gingery, <strong>and</strong> D. Eisenberg.<br />

2005. Amyloid-like fibrils <strong>of</strong> ribonuclease A with threedimensional<br />

domain-swapped <strong>and</strong> native-like structures. Nature.<br />

437:266–269.<br />

83. Ranson, N., T. Stromer, L. Bousset, R. Melki, <strong>and</strong> R. C. Serpell. 2006.<br />

Insights into the architecture <strong>of</strong> the Ure2p yeats protein assemblies from<br />

helical twisted fibrils. Protein Sci. 15:2481–2487.<br />

84. Hatters, D. M., C. A. MacRails, R. Daniels, W. S. Gosal, N. H. Thomson,<br />

et al. 2003. The circularization <strong>of</strong> amyloid fibrils formed by apolipoprotein<br />

C–II. Biophys. J. 85:3979–3990.<br />

85. Lashuel, H. A., <strong>and</strong> P. T. Lansbury Jr. 2006. Are amyloid diseases<br />

caused by protein aggregates that mimic bacterial pore-forming toxins?<br />

Q. Rev. Biophys. 39:167–201.<br />

86. Meehan, S., T. P. J. Knowles, A. J. Baldwin, J. F. Smith, A. M. Squires,<br />

et al. 2007. Characterization <strong>of</strong> amyloid fibril formation by small heatshock<br />

chaperone proteins human aA-, aB- <strong>and</strong> R120G aB-crystallins.<br />

J. Mol. Biol. 372:470–484.<br />

87. Kad, N. M., S. L. Myers, D. P. Smith, S. E. Radford, <strong>and</strong> N. H. Thomson.<br />

2003. Hierarchical assembly <strong>of</strong> b2-microglobulin amyloid in vitro<br />

revealed by atomic force microscopy. J. Mol. Biol. 330:785–797.<br />

88. Khurana, R., J. L. Jimenez, E. J. Nettleton, M. Bouchard, C. V. Robinson,<br />

et al. 2002. The prot<strong>of</strong>ilament structure <strong>of</strong> insulin amyloid fibrils.<br />

Proc. Natl. Acad. Sci. USA. 99:9196–9201.<br />

89. Otzen, D. E., <strong>and</strong> M. Oliveberg. 2004. Transient formation <strong>of</strong> nanocrytalline<br />

structures during fibrillation <strong>of</strong> an Alzheimer-like peptide.<br />

Protein Sci. 13:1417–1421.<br />

90. Grudzielanek, S., V. Smirnovas, <strong>and</strong> R. Winter. 2006. Solvation-assisted<br />

pressure tuning <strong>of</strong> insulin fibrillation: from novel aggregation pathways<br />

to biotechnological applications. J. Mol. Biol. 356:497–506.<br />

91. Green, J. D., C. Goldsbury, J. Kistler, G. K. Cooper, <strong>and</strong> U. Aebi. 2004.<br />

Human amylin oligomer growth <strong>and</strong> fibril elongation define two distinct<br />

phases in amyloid formation. J. Biol. Chem. 279:12206–12212.<br />

92. Chamberlain, A. K., C. E. MacPhee, J. Zurdo, L. A. Morozova-Roche,<br />

H. A. O. Hill, et al. 2000. Ultrastructural organization <strong>of</strong> amyloid fibrils<br />

by atomic force microscopy. Biophys. J. 79:3282–3293.<br />

93. Goldsbury, C. S., S. Wirtz, S. A. Müller, S. Sunderji, P. Wicki, et al.<br />

2000. Studies on the in vitro assembly <strong>of</strong> Aba-40: implications for<br />

the search for Ab fibril formation inhibitors. J. Struct. Biol. 130:<br />

217–231.<br />

94. Petkova, A. T., R. D. Leapman, Z. Guo, W. M. Yau, M. P. Mattson,<br />

et al. 2005. <strong>Self</strong>-propagating, molecular-level polymorphism in<br />

Alzheimer’s b-amyloid fibrils. Science. 307:262–265.<br />

164


Influence <strong>of</strong> Electrostatic Interactions on the Fibrillation Process <strong>of</strong> Human Serum Albumin<br />

Introduction<br />

Josué Juárez, Sonia Goy López, Adriana Cambón, Pablo Taboada,* <strong>and</strong> Víctor Mosquera<br />

Grupo de Física de Coloides y Polímeros, Departamento de Física de la Materia Condensada, Facultad de<br />

Física, UniVersidad de Santiago de Compostela, E-15782, Santiago de Compostela, Spain<br />

ReceiVed: March 12, 2009; ReVised Manuscript ReceiVed: May 7, 2009<br />

The fibrillation propensity <strong>of</strong> the multidomain protein human serum albumin (HSA) has been analyzed under<br />

physiological <strong>and</strong> acidic conditions at room <strong>and</strong> elevated temperatures with varying ionic strengths by different<br />

spectroscopic techniques. The kinetics <strong>of</strong> fibril formation under the different solution conditions <strong>and</strong> the<br />

structures <strong>of</strong> resulting fibrillar aggregates were also determined. In this way, we have observed that fibril<br />

formation is largely affected by electrostatic shielding: at physiological pH, fibrillation is progressively more<br />

efficient <strong>and</strong> faster in the presence <strong>of</strong> up to 50 mM NaCl; meanwhile, at larger salt concentrations, excessive<br />

shielding <strong>and</strong> further enhancement <strong>of</strong> the solution hydrophobicity might involve a change in the energy<br />

l<strong>and</strong>scape <strong>of</strong> the aggregation process, which makes the fibrillation process difficult. In contrast, under acidic<br />

conditions, a continuous progressive enhancement <strong>of</strong> HSA fibrillation is observed as the electrolyte concentration<br />

in solution increases. Both the distinct ionization <strong>and</strong> initial structural states <strong>of</strong> the protein before incubation<br />

may be the origin <strong>of</strong> this behavior. CD, FT-IR, <strong>and</strong> tryptophan fluorescence spectra seem to confirm this<br />

picture by monitoring the structural changes in both protein tertiary <strong>and</strong> secondary structures along the<br />

fibrillation process. On the other h<strong>and</strong>, the fibrillation <strong>of</strong> HSA does not show a lag phase except at pH 3.0 in<br />

the absence <strong>of</strong> added salt. Finally, differences in the structure <strong>of</strong> the intermediates <strong>and</strong> resulting fibrils under<br />

the different conditions are also elucidated by TEM <strong>and</strong> FT-IR.<br />

Protein aggregation <strong>and</strong>, particularly, amyloid formation have<br />

received considerable interest in the fields <strong>of</strong> protein research<br />

<strong>and</strong> clinical medicine, since growing evidence has accumulated<br />

that these processes are likely to be key issues in the etiology<br />

<strong>of</strong> various medical disorders such as Alzheimer’s, Parkinson’s,<br />

Huntington’s, <strong>and</strong> Creutzfeldt-Jakob diseases, type II diabetes,<br />

<strong>and</strong> cystic fibrosis among others. In fact, about 20 different<br />

known syndromes are associated with the formation <strong>of</strong> amyloid<br />

fibrils. 1-6 In addition, a number <strong>of</strong> nondisease associated proteins<br />

have also been found to be able to form ordered cytotoxic<br />

aggregates <strong>and</strong> amyloid fibrils in Vitro. 7 On the other h<strong>and</strong>, the<br />

exceptional physical characteristics <strong>of</strong> the amyloidal protein<br />

state, such as its stability, mechanical strength, <strong>and</strong> resistance<br />

to degradation, imply that this type <strong>of</strong> structure possesses a range<br />

<strong>of</strong> potential technological applications in biotechnology <strong>and</strong><br />

materials science. 8,9<br />

A consensus has emerged that the different processes <strong>of</strong> in<br />

ViVo <strong>and</strong> in Vitro aggregation might be linked by the implication<br />

<strong>of</strong> destabilized protein molecules that undergo non-native<br />

backbone-dominated self-assembly as a competing pathway to<br />

native functional folding. 10 However, both folding <strong>and</strong> protein<br />

aggregation are thought to underlie some common principles<br />

which aim at minimizing the Gibbs energy <strong>of</strong> destabilized<br />

protein states, such as the saturation <strong>of</strong> dangling hydrogen<br />

bonds, 11 the reduction <strong>of</strong> the solvent-accessible surface area, or<br />

the burial <strong>of</strong> hydrophobic residues. 12 In this way, globular<br />

proteins generally need to be destabilized (e.g., by mutation, 13,14<br />

heat, 15,16 high pressure, 17,18 low pH, 19,20 or organic denaturant 7,15 )<br />

to aggregate rapidly, with fibril formation proceeding from<br />

* Author to whom correspondence should be addressed. E-mail:<br />

pablo.taboada@usc.es. Telephone: 0034981563100, ext. 14042. Fax:<br />

0034981520676.<br />

J. Phys. Chem. B 2009, 113, 10521–10529 10521<br />

extensively or partially unfolded states 3,21 or, in some cases,<br />

from native-like states in which unfolding may initially be<br />

limited <strong>and</strong> confined to local regions <strong>of</strong> the protein. 22,23<br />

The physiological importance <strong>of</strong> human serum albumin as a<br />

carrier protein <strong>and</strong> blood pressure regulator <strong>and</strong> its propensity<br />

to easily aggregate in Vitro have become a good model for<br />

protein aggregation studies. As the phenomenon <strong>of</strong> protein<br />

aggregation appears to reflect certain generic “polymeric”<br />

features <strong>of</strong> proteins, 24 the studies <strong>of</strong> the mechanisms <strong>of</strong> protein<br />

aggregation on model systems are extremely useful for a better<br />

underst<strong>and</strong>ing <strong>of</strong> the molecular mechanisms <strong>of</strong> disease-associated<br />

amyloidogenesis. Since in living organisms the native<br />

environment <strong>of</strong> proteins is a complex composition <strong>of</strong> water,<br />

cosolvents, <strong>and</strong> cosolutes which affect the stability <strong>of</strong> the native<br />

protein fold, 25 experimental approaches to mimic various cellular<br />

environments in Vitro to allow investigations on protein folding<br />

<strong>and</strong> aggregation <strong>of</strong>ten utilize the addition <strong>of</strong> cosolvents <strong>and</strong><br />

cosolutes. Thereby, differences in the properties <strong>of</strong> the solvent<br />

might induce solvent-adapted structural “responses” <strong>of</strong> the<br />

aggregating protein. In particular, the importance <strong>of</strong> electrostatic<br />

interactions in the formation <strong>of</strong> amyloid fibrils 20,26,27 <strong>and</strong> the<br />

relevance <strong>of</strong> the total net charge on the protein to the fibril<br />

formation propensity have been pointed out. 26,27<br />

To put more light on this issue, in the present work we present<br />

a systematic study on the relationship between serum albumin<br />

amyloid-like fibrillation kinetics, fibrillation pathways, <strong>and</strong> the<br />

resulting aggregated protein structures under different conditions<br />

where the amyloid fibrils <strong>of</strong> this model protein were found by<br />

changing solution pH <strong>and</strong> solution ionic strength. 28,29 In this<br />

manner, a modulation <strong>of</strong> the electrostatic interactions between<br />

protein molecules <strong>and</strong> aggregates results in different balances<br />

<strong>of</strong> intermolecular forces between protein molecules, which<br />

involve changes in both protein aggregation <strong>and</strong> their kinetics.<br />

In this way, we have found large differences in the fibrillation<br />

10.1021/jp902224d CCC: $40.75 © 2009 American Chemical Society<br />

Published on Web 07/02/2009<br />

165


10522 J. Phys. Chem. B, Vol. 113, No. 30, 2009 Juárez et al.<br />

kinetics <strong>of</strong> HSA <strong>and</strong>, as a result, different structural intermediates<br />

<strong>and</strong> fibrillation pathways. As well as that, differences in<br />

the resulting amyloid-like fibrils under the different solution<br />

conditions were also observed. It seems, then, that both pH <strong>and</strong><br />

ionic strength play a key role in regulating the several steps<br />

which regulate HSA fibrillation.<br />

Experimental Section<br />

Materials. Human serum albumin (70024-90-7), Congo Red<br />

(CR), <strong>and</strong> Thi<strong>of</strong>lavin T (ThT) were obtained from Sigma<br />

Chemical Co <strong>and</strong> used as received. All other chemicals were<br />

<strong>of</strong> the highest purity available.<br />

Preparation <strong>of</strong> HSA Solutions. The buffer solutions used<br />

were glycine + HCl for pH 3.0 <strong>and</strong> sodium monophosphatesodium<br />

diphosphate for pH 7.4, respectively. The buffers were<br />

prepared in different NaCl concentrations (0, 20, 50, 100, 150,<br />

<strong>and</strong> 250 mM NaCl). The HSA was dissolved in each buffer<br />

solution to a final concentration <strong>of</strong> typically 20 mg/mL, if not<br />

otherwise stated. Protein concentration was determined spectrophotometrically<br />

using a molar absorption coefficient <strong>of</strong> 35 219<br />

M-1 cm-1 at 280 nm. 30 Before incubation, the solution was<br />

filtered through a 0.2 µm filter into sterile test tubes. Samples<br />

were incubated at a specified temperature in a refluxed reactor.<br />

Samples were taken out at intervals <strong>and</strong> stored on ice before<br />

adding CR or ThT.<br />

Seeding Solutions. To test whether seeding with preformed<br />

aggregates increases the rate <strong>of</strong> HSA aggregation under the<br />

different conditions where fibrils are formed, a protein solution<br />

was incubated for 24-48 h <strong>and</strong> an aliquot that corresponded to<br />

10% (w/w) <strong>of</strong> the total protein concentration was then added<br />

to a fresh protein solution.<br />

Thi<strong>of</strong>lavin T Spectroscopy. 50 µL aliquots <strong>of</strong> protein<br />

solution were withdrawn at different times <strong>and</strong> diluted 100 times<br />

with 4 mL <strong>of</strong> ThT solution in corresponding buffer (final ThT<br />

concentration 20 µM). Samples were continuously stirred during<br />

measurements. Fluorescence was measured in a Cary Eclipse<br />

fluorescence spectrophotometer equipped with a temperature<br />

control device <strong>and</strong> a multicell sample holder (Varian Instruments<br />

Inc.). Excitation <strong>and</strong> emission wavelengths were 450 <strong>and</strong> 482<br />

nm, respectively. All intensities were background-corrected for<br />

the ThT-fluorescence in the respective buffer without the protein.<br />

Congo Red Binding. Changes in the absorbance <strong>of</strong> Congo<br />

Red dye produced by binding to HSA were measured in a<br />

UV-vis spectrophotometer DU Series 640 (Beckman Instruments)<br />

operating from 190 to 1100 nm. All measurements were<br />

made in the wavelength range 200-600 nm in matched quartz<br />

cuvettes. Protein solutions were diluted 20-200-fold into a<br />

buffer solution with 5 µM Congo Red (Acros Organics, Geel,<br />

Belgium). Spectra in the presence <strong>of</strong> the dye were compared<br />

with that <strong>of</strong> the buffer containing Congo Red in the absence <strong>of</strong><br />

protein, <strong>and</strong> also with that containing protein without dye.<br />

Circular Dichroism (CD). Far-UV circular dichroism (CD)<br />

spectra were obtained using a JASCO-715 automatic recording<br />

spectropolarimeter with a JASCO PTC-343 Peltier-type thermostatted<br />

cell holder. Quartz cuvettes with 0.2 cm path length<br />

were used. CD spectra were obtained from aliquots withdrawn<br />

from the aggregation mixtures at the indicated conditions <strong>and</strong><br />

incubation times, <strong>and</strong> recorded between 195 <strong>and</strong> 250 nm at 25<br />

°C. The mean residue ellipticity θ (deg cm2 dmol-1 ) was<br />

calculated from the formula θ ) (θobs/10)(MRM/lc), where θobs<br />

is the observed ellipticity in deg, MRM is the mean residue<br />

molecular mass, l is the optical path-length (in cm), <strong>and</strong> c is<br />

the protein concentration (in g mL-1 ). In order to calculate the<br />

composition <strong>of</strong> the secondary structure <strong>of</strong> the protein, SEL-<br />

CON3, CONTIN, <strong>and</strong> DSST programs were used to analyze<br />

far-UV CD spectra. Final results were assumed when data<br />

generated from all programs show convergence. 31<br />

Fourier Transform Infrared Spectroscopy (FT-IR). FT-<br />

IR spectra <strong>of</strong> HSA in aqueous solutions were determined by<br />

using a FT-IR spectrometer (model IFS-66v from Bruker)<br />

equipped with a horizontal ZnS ATR accessory. The spectra<br />

were obtained at a resolution <strong>of</strong> 2 cm -1 , <strong>and</strong> generally 200 scans<br />

were accumulated to get a reasonable signal-to-noise ratio.<br />

Solvent spectra were also examined in the same accessory <strong>and</strong><br />

at the same instrument conditions. Each different sample<br />

spectrum was obtained by digitally subtracting the solvent<br />

spectrum from the corresponding sample spectrum. Each sample<br />

solution was repeated three times to ensure reproducibility <strong>and</strong><br />

was averaged to produce a single spectrum.<br />

Protein Fluorescence. To examine the conformational variations<br />

around the tryptophan residue <strong>of</strong> HSA, fluorescence<br />

emission spectra <strong>of</strong> HSA samples were excited at 295 nm, which<br />

provides no excitation <strong>of</strong> tyrosine residues, <strong>and</strong> therefore, neither<br />

emission nor energy transfer to the lone side chain takes place.<br />

Slit widths were typically 5 nm. The spectrophotometer used<br />

was a Cary Eclipse from Varian Instruments Inc.<br />

Transmission Electron Microscopy (TEM). Suspensions <strong>of</strong><br />

HSA were applied to carbon-coated copper grids, blotted,<br />

washed, negatively stained with 2% (w/v) <strong>of</strong> phosphotungstic<br />

acid, air-dried, <strong>and</strong> then examined with a Phillips CM-12<br />

transmission electron microscope operating at an accelerating<br />

voltage <strong>of</strong> 120 kV. Samples were diluted between 20 <strong>and</strong> 200fold<br />

where prior deposition on the grids was needed.<br />

X-ray Diffraction. X-ray diffraction experiments were carried<br />

out using a Siemens D5005 rotating anode X-ray generator.<br />

Twin Göbel mirrors were used to produce a well-collimated<br />

beam <strong>of</strong> Cu KR radiation (λ ) 1.5418 A). Samples were put<br />

into capillaries with diameters <strong>of</strong> 0.5 mm. X-rays diffraction<br />

patterns were recorded with an AXS F.Nr. J2-394 imaging plate<br />

detector.<br />

Results <strong>and</strong> Discussion<br />

It has been previously reported that amyloid fibrils can also<br />

be formed in Vitro from R-globular proteins such as myoglobin,<br />

ovoalbumin,bovine(BSA),<strong>and</strong>human(HSA)serumalbumins, 4,32,33<br />

for example. Similar to these proteins, native HSA also lacks<br />

any properties suggesting a predisposition to form amyloid<br />

fibrils, since most <strong>of</strong> its sequence (>60%) is arranged in an<br />

R-helix structure, with the subsequent tightening <strong>of</strong> its structure<br />

through intramolecular interactions such as hydrogen bonds. In<br />

this way, serum albumin aggregation is only promoted under<br />

conditions favoring partly destabilized monomers <strong>and</strong> dimers<br />

such as low pH, high temperatures, or the presence <strong>of</strong> chemical<br />

denaturants. 34 In a recent report, 28 we have shown that partially<br />

destabilized HSA molecules form amyloid-like fibrils <strong>and</strong> other<br />

types <strong>of</strong> aggregates under different solution conditions such as<br />

increasing the solution temperature <strong>and</strong> changing the solution<br />

pH. 28,35,36 Thus, HSA was subjected to conditions previously<br />

found to be effective for protein aggregation, in particular those<br />

inducing amyloid-like fibril formation but modifying the solution<br />

ionic strength in order to analyze its effects on the amyloidlike<br />

assembly pathway <strong>of</strong> this protein. Thus, we prepared HSA<br />

solutions at a concentration <strong>of</strong> 20 mg/mL in 0.01 M sodium<br />

phosphate buffer (pH 7.4) or 0.01 M glycine-HCl buffer (pH<br />

3.0) in the presence <strong>of</strong> 0, 20, 50, 100, 150, <strong>and</strong> 250 mM NaCl<br />

at 25 or 65 °C followed by subsequent incubation for up to 15<br />

days. HSA is in its native form (N) at pH 7.4, which is composed<br />

mostly <strong>of</strong> R-helix secondary structure. In contrast, HSA adopts<br />

166


Fibrillation Process <strong>of</strong> Human Serum Albumin J. Phys. Chem. B, Vol. 113, No. 30, 2009 10523<br />

Figure 1. Time evolution <strong>of</strong> ThT fluorescence in HSA solutions<br />

incubated at 65 °C at (a) pH 7.4 <strong>and</strong> (b) pH 3.0 in the presence <strong>of</strong> (9)<br />

0, (O) 20, (2) 50, (3) 100, ([) 150, <strong>and</strong> (+) 250 mM NaCl.<br />

an extended less-ordered configuration (E-state) below pH 3.5<br />

with lesser helical content. 37-39 On the other h<strong>and</strong>, 65 °C was<br />

chosen as the incubation temperature, provided that HSA<br />

temperature-induced denaturation takes place through a twostate<br />

transition with a first melting temperature, Tm, <strong>of</strong>∼56 °C<br />

<strong>and</strong> a second Tm <strong>of</strong> ∼62 °C 35,40,41 as a consequence <strong>of</strong> the<br />

sequential unfolding <strong>of</strong> the different domains <strong>of</strong> the protein, in<br />

particular IIA <strong>and</strong> IIIA subdomains: at 50 °C, a reversible<br />

separation <strong>of</strong> site I <strong>and</strong> II occurs; below 70 °C, the irreversible<br />

unfolding <strong>of</strong> site II is present, while increasing the temperature<br />

over 70 °C or higher results in the irreversible unfolding<br />

<strong>of</strong> site I. 42 Moreover, as the pH becomes more acidic, Tm<br />

becomes lower. 35<br />

Amyloid Formation Kinetics. When HSA solutions (20 mg/<br />

mL in the presence <strong>of</strong> 0, 20, 50, 100, 150, or 250 mM NaCl)<br />

were incubated at room temperature up to 300 h, the ThT<br />

fluorescence emission intensity was negligible at pH 7.4 whereas<br />

a slight increase was observed at pH 3.0 as a consequence <strong>of</strong> a<br />

little gain in -sheet structure due to the formation <strong>of</strong> some<br />

oligomeric aggregates (data not shown). This suggests that HSA<br />

was not capable <strong>of</strong> forming amyloid fibrils or other types <strong>of</strong><br />

aggregates with significant -sheet structure under these conditions.<br />

In contrast, a time-dependent increase in ThT fluorescence<br />

is observed when the different HSA solutions were incubated<br />

at 65 °C (see Figure 1), in accordance with fibril formation.<br />

Congo Red (CR) absorption was also used to corroborate the<br />

formation <strong>of</strong> these fibrils in HSA solutions, which display a<br />

red-shift <strong>of</strong> the differential absorption maximum from 495 to<br />

ca. 530 nm at 65 °C (see Figure S1 as an example).<br />

In general, the kinetics <strong>of</strong> HSA fibrillation under the present<br />

solution conditions at 65 °C involved the continuous rising <strong>of</strong><br />

TABLE 1: Kinetic Parameters <strong>of</strong> the <strong>Self</strong>-<strong>Assembly</strong> Process<br />

<strong>of</strong> HSA Solutions<br />

pH NaCl (mM) ∆F ksp (h-1 ) n<br />

7.4 0 83 0.019 0.56<br />

20 82 0.040 0.70<br />

50 83 0.084 0.92<br />

100 57 0.068 0.80<br />

150 52 0.056 0.77<br />

250 46 0.034 0.70<br />

3.0 0 6/22 0.144/0.005 1.25/3.60<br />

20 44 0.008 1.25<br />

50 57 0.010 0.96<br />

100 73 0.012 0.95<br />

150 77 0.024 0.93<br />

250 56 0.233 1.10<br />

ThT fluorescence during the early periods <strong>of</strong> incubation until a<br />

quasi-plateau region was attained in the time scale analyzed,<br />

as occurred for different proteins such as BSA, 32 acylphosphatase,<br />

7 or ovoalbumin, 33 <strong>and</strong> it exhibited no discernible lag<br />

phase. This absence may be related to the fact that the initial<br />

protein aggregation is a downhill process, which does not require<br />

a highly organized <strong>and</strong> stable nucleus. This was confirmed by<br />

the absence <strong>of</strong> any remarkable effect on the fluorescence curves<br />

when protein seeds are added to protein solutions followed by<br />

subsequent incubation (see Figure S2 in the Supporting Information).<br />

It has been suggested that large multidomain proteins<br />

such as BSA are able to form propagation-competent nucleuslike<br />

structures (oligomeric structures), since there is no energy<br />

barrier to impede aggregate growth. 32 As has been discussed in<br />

detail previously before, 32,36 HSA forms these oligomers upon<br />

very short incubation times, which usually occurs by a mechanism<br />

<strong>of</strong> classical coagulation or downhill polymerization. 44<br />

Nevertheless, a different behavior was found at acidic pH in<br />

the absence <strong>of</strong> added electrolyte: After a small increase in ThT<br />

fluorescence at very short incubation times, there exists an<br />

almost plateau region (between 24 <strong>and</strong> 100 h approximately),<br />

from which the ThT fluorescence starts to increase again (see<br />

Figure 1b). This plateau region was identified as a lag phase,<br />

as confirmed by a fibril seeding growth process (see Figure S3<br />

in the Supporting Information). This might well indicate that<br />

the formation <strong>of</strong> oligomeric structures might need more time<br />

to be developed <strong>and</strong>/or persist for longer times due to their<br />

enhanced solubility under acidic conditions in the absence <strong>of</strong><br />

added salt, so oligomers need to reach a certain number/size to<br />

change the energy l<strong>and</strong>scape <strong>of</strong> the system <strong>and</strong> promote further<br />

aggregation.<br />

Figure 1 also shows that amyloid formation was favored <strong>and</strong><br />

took place progressively quicker as the electrolyte concentration<br />

increased from 0 to 50 mM NaCl at pH 7.4 <strong>and</strong> at the whole<br />

concentration range studied at pH 3.0, as denoted by the<br />

increment in ThT fluorescence intensity <strong>and</strong> the subsequent<br />

presence <strong>of</strong> a plateau region at shorter incubation times. This<br />

points out that the increasing hydrophobicity originated from<br />

the screening <strong>of</strong> repulsive electrostatic interactions between<br />

protein molecules is capable <strong>of</strong> stabilizing the -sheets. 36,45-48<br />

In contrast, a progressive decrease in ThT fluorescence intensity<br />

is observed at electrolyte concentrations larger than 50 mM at<br />

physiological pH, with the plateau located at shorter incubation<br />

times. This decrease can be related to the formation <strong>of</strong> bundles<br />

<strong>of</strong> shorter fibrils <strong>and</strong> to the presence <strong>of</strong> amorphous aggregates<br />

(with larger R-helix <strong>and</strong> unordered conformation contents, as<br />

will be shown below), <strong>and</strong> is a consequence <strong>of</strong> an excessive<br />

shielding <strong>of</strong> intermolecular electrostatic forces, which is accompanied<br />

by an additional enhancement <strong>of</strong> the solution<br />

167


10524 J. Phys. Chem. B, Vol. 113, No. 30, 2009 Juárez et al.<br />

Figure 2. Far-UV spectra <strong>of</strong> HSA solutions at 65 °C <strong>and</strong> pH 7.4 in the presence <strong>of</strong> (a) 0, (b) 50, (c) 100, <strong>and</strong> (d) 150 mM NaCl at (1) 0 h, (2)<br />

12 h, (3) 24 h, (4) 48, (5) 60 h, <strong>and</strong> (6) 72 h <strong>of</strong> incubation.<br />

hydrophobicity through (a) a side-to-side adsorption between<br />

protein molecules <strong>and</strong> fibrils; (b) a reduction <strong>of</strong> the ThT<br />

fluorescence by blocking ThT binding sites on the fibrils due<br />

to a reduction in exposed surface area; 49 <strong>and</strong> (c) a lower fraction<br />

<strong>of</strong> formed fibrils due to an unsuitable energy balance to establish<br />

effective intermolecular interactions; that is, r<strong>and</strong>om aggregation<br />

is promoted, so there exists a change in the energy l<strong>and</strong>scape<br />

<strong>of</strong> the protein fibrillation process. 50<br />

Differences in the trends observed between physiological <strong>and</strong><br />

acidic conditions in the presence <strong>of</strong> electrolyte might arise from<br />

the fact that ordered aggregation <strong>of</strong> globular proteins requires<br />

partial unfolding <strong>of</strong> the native state, 21 which enables this<br />

precursor state to be populated <strong>and</strong> to expose aggregationcompetent<br />

regions that are usually protected against intermolecular<br />

interactions in the native protein. However, these regions<br />

can be different <strong>and</strong> display different propensities to solvent<br />

exposition depending on solution conditions. Thus, although<br />

amyloid formation can occur, rates <strong>and</strong> extents <strong>of</strong> fibril<br />

formation, fibrillation pathways, <strong>and</strong> the structures <strong>of</strong> the<br />

resulting fibrils can be very diverse. 51,52 In this way, it seems<br />

that both the distinct ionization state <strong>of</strong> the protein <strong>and</strong> the<br />

different initial structural state at acidic conditions (exp<strong>and</strong>ed-<br />

E-state) may be the origin <strong>of</strong> the observed differences, which<br />

change the energy l<strong>and</strong>scape <strong>of</strong> the aggregation process.<br />

The time-dependence <strong>of</strong> the ThT fluorescence curves was<br />

fitted by nonlinear square curve-fitting to a stretched exponential<br />

function F ) F ∞ + ∆F exp(-[kspt] n ) in order to derive<br />

information on the amyloid kinetics formation. F, F ∞ , <strong>and</strong> ∆F<br />

are the observed fluorescence intensity at time t, the final<br />

fluorescence intensity, <strong>and</strong> the fluorescence amplitude, respectively,<br />

<strong>and</strong> ksp is the rate <strong>of</strong> spontaneous fibril formation. Values<br />

<strong>of</strong> ksp, n, <strong>and</strong> ∆F are shown in Table 1. Values <strong>of</strong> n lower than<br />

1 indicated that the kinetics can be approximated to several<br />

exponential functions indicative <strong>of</strong> the existence <strong>of</strong> multiple<br />

events in the amyloid formation. The increasing ksp values as<br />

the electrolyte concentration goes up to 100 mM at physiological<br />

pH <strong>and</strong> at the whole concentration range under acidic conditions<br />

corroborate the more efficient aggregation due to electrostatic<br />

screening in the presence <strong>of</strong> added salt. However, under<br />

physiological conditions, progressively lower ksp <strong>and</strong> n values<br />

are derived at larger ionic strengths (see Table 1), maybe as a<br />

predominance <strong>of</strong> fibril bundling <strong>and</strong> formation <strong>of</strong> amorphous<br />

aggregates over that <strong>of</strong> new fibrils, as commented before. This<br />

behavior is in contrast to that observed for other albumin<br />

proteins such as ovoalbumin <strong>and</strong> BSA, for which their fibrillation<br />

mechanisms <strong>and</strong>, thus, their aggregation rates were found<br />

not to be so strongly influenced by ionic strength if compared<br />

to that <strong>of</strong> HSA. 33 In particular, for ovolabumin, it was found<br />

that a continuous slight decrease in the rate <strong>of</strong> fibrillation takes<br />

place as the electrolyte concentration occurs, whereas for BSA<br />

the trend is similar to that <strong>of</strong> HSA, although less marked.<br />

Possibly, the subtle differences in the amino acid sequence<br />

between the proteins <strong>and</strong> in their structural conformation may<br />

account for these differences.<br />

On the other h<strong>and</strong>, it is interesting to note that, after 3 days<br />

<strong>of</strong> incubation, most HSA solutions transformed into a gel-like<br />

state at elevated ionic strengths (between 100 <strong>and</strong> 250 mM). It<br />

is known that salt concentrations have an important role, not<br />

only favoring conformational transitions <strong>and</strong> intramolecular<br />

interactions in the protein but also enhancing protein-protein<br />

interactions leading to gelification. 48-50 In our case, these gels<br />

can be a consequence <strong>of</strong> a “correct” balance between electrostatic<br />

shielding <strong>and</strong> hydrophobic interactions which enables<br />

fibrils to interact in suitable relative orientations, as has been<br />

previously demonstrated for other proteins. 50 Nevertheless, these<br />

effects will be discussed in a forthcoming publication.<br />

168


Fibrillation Process <strong>of</strong> Human Serum Albumin J. Phys. Chem. B, Vol. 113, No. 30, 2009 10525<br />

Figure 3. Far-UV spectra <strong>of</strong> HSA solutions at 65 °C <strong>and</strong> pH 3.0 in the presence <strong>of</strong> (a) 0, (b) 50, <strong>and</strong> (c) 150 mM NaCl at (1) 0 h, (2) 24 h, (3)<br />

50 h, (4) 100 h, (5) 150 h, <strong>and</strong> (6) 250 h <strong>of</strong> incubation.<br />

TABLE 2: Secondary Structure Variations (in %) on HSA<br />

Samples upon Incubation at 65° C under Different Solution<br />

pH Values <strong>and</strong> Electrolyte Concentrations<br />

pH NaCl (mM) ∆R-helix ∆-sheet ∆turn ∆ unordered<br />

7.4 0 -35 16 8 11<br />

20 -37 18 9 12<br />

50 -42 21 10 11<br />

100 -39 16 9 14<br />

150 -37 15 8 16<br />

250 -33 11 7 19<br />

3.0 0 -20 9 3 8<br />

20 -25 12 4 9<br />

50 -28 14 4 10<br />

100 -32 17 5 10<br />

150 -36 19 6 11<br />

250 -35 20 5 10<br />

Structural Changes. To shed additional light on the differences<br />

<strong>of</strong> the amyloid-like formation <strong>of</strong> HSA in the presence <strong>of</strong><br />

added salt at the different pH values, we recorded CD, FT-IR,<br />

<strong>and</strong> tryptophan (Trp) fluorescence spectra. The far-UV CD<br />

spectra <strong>of</strong> HSA solutions incubated at room temperature in the<br />

absence <strong>and</strong> the presence <strong>of</strong> electrolyte showed two minima at<br />

208 <strong>and</strong> 222 nm, characteristic <strong>of</strong> a helical structure. At acidic<br />

pH, HSA molecules also preserve an important part <strong>of</strong> their<br />

R-helix structure upon incubation, as denoted by the presence<br />

<strong>of</strong> the minima at 208 <strong>and</strong> 222 nm. However, a slight decrease<br />

in ellipticity, [θ], is observed, which is compatible with a little<br />

decrease <strong>of</strong> R-helices due to the appearance <strong>of</strong> small amorphous<br />

aggregates in solution (not shown).<br />

When HSA solutions were incubated at 65 °C, far-UV CD<br />

spectra showed a dramatic change in ellipticity throughout<br />

incubation. Figures 2 <strong>and</strong> 3 show CD spectra for HSA solutions<br />

at 65 °C in the presence <strong>of</strong> several NaCl concentrations at both<br />

pH 7.4 <strong>and</strong> 3.0, respectively. Under these conditions, the<br />

minimum at 222 nm progressively disappears while the negative<br />

b<strong>and</strong> intensity at 208 nm is also reduced. All this indicates that<br />

high-temperature conditions spawn intermediates, which are<br />

clearly less helical than the starting conformations (a decrease<br />

in R-helix content is observed). Thus, these changes obtained<br />

in CD spectra suggest the increments <strong>of</strong> -sheet <strong>and</strong> loop<br />

structures at longer incubation times (∼72 h <strong>and</strong> ∼150 h for<br />

pH 7.4 <strong>and</strong> 3.0, respectively), which resemble those typical <strong>of</strong><br />

proteins with important proportions <strong>of</strong> -sheet structure characterized<br />

by a minimum at ca. 215-220 nm <strong>and</strong> small [θ] values<br />

as a consequence <strong>of</strong> the presence <strong>of</strong> -rich aggregates in<br />

solution, as can be observed, for example, in Figure 2b.<br />

Nevertheless, it is worth mentioning that the increasing number<br />

<strong>and</strong> size <strong>of</strong> large scattering objects such as fibrils in solution as<br />

the incubation proceeds involves the strong decrease in ellipticity,<br />

which makes it a little less clear the ellipticity minimum<br />

around 215-220 nm corresponding to -sheet structures.<br />

Changes in protein structure composition derived at the end<br />

<strong>of</strong> incubation at 65 °C were obtained from CD analysis 31 <strong>and</strong><br />

are shown in Table 2. Data in this table are displayed as the<br />

difference (expressed in %) between the beginning <strong>and</strong> the end<br />

<strong>of</strong> the incubation process <strong>of</strong> protein structure composition (∆Rhelix,<br />

∆-sheet, ∆turn, <strong>and</strong> ∆r<strong>and</strong>om coil conformation, respectively).<br />

The negative signs mean a decrease in the corresponding<br />

structure. At pH 7.4 <strong>and</strong> room temperature, the initial<br />

R-helix content is ∼59%, the -sheet conformation is ∼5%,<br />

the turn content is ∼13%, <strong>and</strong> the remaining r<strong>and</strong>om coil content<br />

is ∼23%, in agreement with previous reports. 47 When HSA<br />

solutions were incubated at 65 °C, these showed important<br />

differences in the secondary structure compositions (see Table<br />

2). The largest changes were observed in R-helix <strong>and</strong> -sheet<br />

169


10526 J. Phys. Chem. B, Vol. 113, No. 30, 2009 Juárez et al.<br />

Figure 4. Second derivative <strong>of</strong> FT-IR spectra at 65 °C <strong>and</strong> (a) pH 7.4 <strong>and</strong> (b) pH 3.0 in the presence <strong>of</strong> (1) 0, (2) 20, (3) 50, (4) 100, (5) 150, <strong>and</strong><br />

(6) 250 mM NaCl after incubation.<br />

Figure 5. Time evolution <strong>of</strong> Trp fluorescence <strong>of</strong> HSA solutions upon<br />

incubation at (a) pH 7.4 <strong>and</strong> (b) pH 3.0 in the presence <strong>of</strong> (9) 0,(O)<br />

20, (∆) 50, (1) 100, <strong>and</strong> (]) 150 mM, <strong>and</strong> (b) at pH 3.0 <strong>and</strong> 25 °C in<br />

the presence <strong>of</strong> 50 mM NaCl.<br />

contents, with the former mainly decreased to be converted into<br />

the -sheet structure. Moreover, coil <strong>and</strong> turn conformations<br />

also change throughout the incubation process, increasing<br />

slightly their significance.<br />

On the other h<strong>and</strong>, increases in the -sheet content become<br />

more important as the electrolyte concentration increases, but<br />

only up to 50 mM NaCl at physiological conditions <strong>and</strong> in the<br />

whole concentration range at acidic pH, for which these<br />

variations were lighter. In this regard, fully mature amyloid<br />

fibrils could not be formed at low electrolyte concentrations at<br />

acidic pH, provided that effective interactions between -str<strong>and</strong>s<br />

are not allowed due to electrostatic repulsions: recent studies<br />

using hydrogen exchange <strong>and</strong> proteolysis have indicated a less-<br />

ordered structure for prot<strong>of</strong>ilaments compared with mature<br />

amyloid fibrils, 54,55 in agreement with their lower -sheet content<br />

suggested by CD <strong>and</strong> ThT binding studies, as in the present<br />

study. On the other h<strong>and</strong>, certain recovery <strong>of</strong> R-helical content<br />

is observed at pH 7.4 at 150 <strong>and</strong> 250 mM NaCl concentrations,<br />

which agrees with the decrease in ThT fluorescence intensity<br />

observed at these concentrations, as commented previously.<br />

FT-IR spectra recorded for HSA solutions under different<br />

conditions at the beginning <strong>and</strong> the end <strong>of</strong> the incubation process<br />

additionally corroborated ThT <strong>and</strong> CD data. Figure 4 shows<br />

the second derivative <strong>of</strong> the HSA solutions at 65 °C atpH7.4<br />

<strong>and</strong> pH 3.0 for different electrolyte concentrations. Following<br />

incubation at 65 °C, a red-shift from 1652 to 1658 cm -1 (see<br />

Figure S4 in the Supporting Information for the FT-IR spectrum<br />

<strong>of</strong> native HSA) <strong>and</strong> a decrease in intensity <strong>of</strong> the amide I b<strong>and</strong><br />

occur up to 50 mM NaCl at physiological conditions, which<br />

are indicative <strong>of</strong> the presence <strong>of</strong> a certain increase <strong>of</strong> disordered<br />

structure, as also seen by far-UV CD. For pH 3.0, this shift is<br />

more progressive along the whole added salt concentration<br />

range. The appearance <strong>of</strong> a well-defined peak around 1625 cm -1<br />

points out a structural transformation to an intermolecular<br />

hydrogen-bonded--sheet structure, 56 which is a structural<br />

characteristic <strong>of</strong> the amyloid fibrils <strong>and</strong> is present at both pH<br />

values. The spectra also show a high frequency component<br />

(∼1693 cm -1 ) at these electrolyte concentrations that would<br />

suggest the existence <strong>of</strong> a -sheet in an antiparallel configuration.<br />

57 The decrease <strong>of</strong> the peak intensity at 1657 cm -1 <strong>and</strong> the<br />

increase at 1625 cm -1 as the electrolyte concentration increases<br />

from 0 to 50 mM NaCl at pH 7.4 support that the major<br />

structural transition is from R-helix f -sheet <strong>and</strong> that<br />

intermolecular interactions are favored at low <strong>and</strong> medium salt<br />

concentrations, in agreement with CD data. In contrast, the<br />

decrease in the b<strong>and</strong> at 1625 cm -1 <strong>and</strong> the blue shift <strong>of</strong> the<br />

b<strong>and</strong> at 1542 cm -1 (not shown) at larger electrolyte concentrations<br />

at this pH point out a lower -sheet configuration <strong>and</strong> a<br />

further unordered conformation, which confirms that large<br />

electrolyte concentrations do not facilitate amyloid formation,<br />

in agreement with CD <strong>and</strong> ThT fluorescence data.<br />

Tryptophanyl fluorescence data showed that when temperature<br />

is raised to 65 °C, a reduction in fluorescence intensity <strong>and</strong> a<br />

hypsochromic shift were observed, as noted in detail previously.<br />

36 Nevertheless, we found that when incubation proceeds<br />

at electrolyte concentrations larger than 100 mM NaCl at pH<br />

7.4, a slightly lower decrease in fluorescence intensity takes<br />

place, which agrees with the assumption that less ordered<br />

amyloid-like fibrils are formed under these solution conditions<br />

in coexistence with amorphous aggregates. In contrast, at acidic<br />

pH, the decrease in fluorescence is also slightly larger as the<br />

170


Fibrillation Process <strong>of</strong> Human Serum Albumin J. Phys. Chem. B, Vol. 113, No. 30, 2009 10527<br />

Figure 6. TEM pictures <strong>of</strong> the different stages <strong>of</strong> the HSA fibrillation process at 25 °C at pH 7.4: (a) in the presence <strong>of</strong> 50 mM NaCl after 150 h<br />

<strong>of</strong> incubation; (b) at pH 3.0 after 150 h <strong>of</strong> incubation in the absence <strong>of</strong> added salt; (c) at pH 7.4 at 65 °C in the presence <strong>of</strong> 150 mM NaCl after<br />

50 h <strong>of</strong> incubation; (d) at pH 7.4 at 65 °C in the absence <strong>of</strong> added NaCl after 50 h <strong>of</strong> incubation; (e) at 65 °C at acidic pH in the presence <strong>of</strong> 100<br />

mM after 150 h <strong>of</strong> incubation; (f) at 65 °C at acidic pH in the presence <strong>of</strong> 250 mM after 50 h <strong>of</strong> incubation.<br />

added salt concentration increases, which denotes more important<br />

changes in tertiary structure (see Figure 5).<br />

In Situ Observation <strong>of</strong> Protein Structures. TEM images<br />

were recorded at different stages during incubation under the<br />

different conditions described above to denote the possible<br />

differences existing between the aggregates as denoted previously.<br />

Distinct time-dependent morphological stages can be<br />

observed in these images. At room temperature, neither fibrils<br />

nor other type <strong>of</strong> aggregates are detected at the beginning <strong>of</strong><br />

incubation. Only after some hours (24 h at pH 7.4, <strong>and</strong> 48 h at<br />

pH 3.0, respectively) are some spherical aggregates detected.<br />

These developed to amorphous clusters, which possess largely<br />

helical structure in the presence <strong>of</strong> electrolyte upon longer<br />

incubation (150 h), as seen in Figure 6a. In addition, at acidic<br />

pH, less numerous, more elongated aggregates can also be<br />

observed, with sizes <strong>of</strong> 20-30 nm in length <strong>and</strong> 3-4 nmin<br />

width, <strong>and</strong> their population slightly becomes more important<br />

as incubation proceeds <strong>and</strong> electrolyte concentration becomes<br />

larger (Figure 6b). The formation <strong>of</strong> these types <strong>of</strong> aggregates<br />

is characterized by a decrease <strong>of</strong> helical structure accompanied<br />

with a slight rise in -sheet <strong>and</strong> unordered conformations at<br />

long incubation times, as seen from CD measurements.<br />

In contrast, when temperature is raised to 65 °C, fibril<br />

formation is observed. Electron microscopy has pointed out that<br />

aggregation leads first to globular species, which subsequently<br />

convert to fibrils with a curly morphology via several structural<br />

intermediates, as has been previously discussed in detail. 36 X-ray<br />

diffraction data also confirms fibril formation (see Figure 7).<br />

Two strong reflections at 4.8 Å <strong>and</strong> 11 Å are observed in the<br />

X-ray image. The 4.8 Å meridional reflection arises from the<br />

spacing between hydrogen-bonded individual str<strong>and</strong>s in the sheet<br />

structure that lie perpendicular to the fibril axis, <strong>and</strong> the<br />

11 Å equatorial reflection corresponds to the intersheet spacing,<br />

with the -sheets stacked face-to face to form the core structure<br />

171


10528 J. Phys. Chem. B, Vol. 113, No. 30, 2009 Juárez et al.<br />

Figure 7. X-ray diffraction pattern <strong>of</strong> HSA fibrils at pH 7.4 <strong>and</strong><br />

65 °C.<br />

<strong>of</strong> prot<strong>of</strong>ilaments. 58,59 This indicates that fibrils possess a cross-<br />

structure, one <strong>of</strong> the diagnostic halfmarks <strong>of</strong> amyloid structures.<br />

Nevertheless, the presence <strong>of</strong> larger concentrations <strong>of</strong> added<br />

electrolyte induces some changes in the structure <strong>of</strong> the resulting<br />

aggregates. In this way, at salt concentrations larger than 100<br />

mM, we have detected that the resulting fibrils appear to be<br />

shorter <strong>and</strong> thicker, <strong>and</strong> coexist with amorphous material (see<br />

Figure 6c <strong>and</strong> d for comparison). Enhanced attractive interactions<br />

between structural intermediates such as beadlike structures<br />

<strong>and</strong> lateral prot<strong>of</strong>ibril previously observed can be the origin <strong>of</strong><br />

these differences. Since electrolyte shields electrostatic repulsion<br />

between charged groups in the protein <strong>and</strong> between protein<br />

chains, the hydrophobic effect is progressively more important.<br />

Electrostatic shielding allows fibril bonding not only between<br />

fibril ends but also between axial sides <strong>of</strong> adjacent prot<strong>of</strong>ibrils<br />

<strong>and</strong> protein molecules, resulting in short <strong>and</strong> thicker fibrils. In<br />

particular, in the case when electrostatic shielding is so effective<br />

that protein association occurs before getting suitable relative<br />

orientations to give regular fibril assemblies, nonspecific aggregation<br />

may take place, leading to the formation <strong>of</strong> amorphous<br />

aggregates, provided that a suitable balance between interactions<br />

is necessary to completely stabilize the -sheet structure.<br />

On the other h<strong>and</strong>, at acidic conditions, the presence <strong>of</strong><br />

increasing electrolyte concentrations allows further growth <strong>of</strong><br />

the short prot<strong>of</strong>ibrils presented in solution at 48-96h<strong>of</strong>incubation<br />

(depending on conditions), which do not develop<br />

otherwise. The corresponding CD spectrum shows a progressive<br />

increase in the proportion <strong>of</strong> the -structure at the expense <strong>of</strong><br />

the helical one during incubation. The presence <strong>of</strong> electrolyte<br />

concentrations larger than 50 mM involves other structural<br />

intermediates in the self-assembly process as ring-shaped<br />

structures, as seen before (see also Figure S5 in the Supporting<br />

Information). 36,60 Furthermore, as the added electrolyte concentration<br />

increases, there is an increase in size <strong>and</strong> a denser<br />

population <strong>of</strong> fibrils, in contrast to the behavior observed for<br />

physiological conditions (see Figure 6f): From the previously<br />

existing fibrils at low salt concentrations possessing sizes ranging<br />

from ∼100 nm to several hundred nanometers in length <strong>and</strong><br />

8-11 nm in width (shorter on average than those obtained at<br />

physiological conditions), which are composed <strong>of</strong> a seam <strong>of</strong><br />

-sheet structure decorated by relatively disorganized R-helical<br />

structure as occurred for Rnase 61 or yeast Ure2p, 62 it is possible<br />

to obtain larger <strong>and</strong> thicker curly fibrils at high electrolyte<br />

concentrations due to electrostatic screening, quite similar to<br />

those observed at pH 7 (Figure 6g).<br />

Thus, it is interesting to note that a similar curly morphology<br />

for HSA fibrils can be achieved under the different solution<br />

conditions. Nevertheless, the rate <strong>and</strong> extent <strong>of</strong> aggregation<br />

depends on solution conditions. On the other h<strong>and</strong>, it is worth<br />

mentioning that, in contrast to HSA, BSA fibrillation has been<br />

shown to be halted at the early curly stage in spite <strong>of</strong> the<br />

enormous structural similarities with HSA, since no further<br />

development in fibrillar structure over long incubation times<br />

has been observed. 35 On the other h<strong>and</strong>, it was also noted that<br />

the formation <strong>of</strong> the mature fibrils in vitro is concomitant with<br />

gelation, as occurred in the present study. 63-65<br />

Conclusions<br />

HSA has the ability to assemble into amyloid-like aggregates<br />

under different solution conditions, i.e., under both physiological<br />

<strong>and</strong> acidic pH at elevated temperature <strong>and</strong> different ionic<br />

strengths. However, the fibrillation process is strongly influenced<br />

by the latter parameter. In this way, we have analyzed the HSA<br />

fibrillation process <strong>and</strong> its kinetics by ThT fluorescence spectroscopy.<br />

We observed that, at physiological pH, fibrillation is<br />

progressively faster <strong>and</strong> more efficient in the presence <strong>of</strong> up to<br />

50 mM NaCl due to electrostatic shielding. However, at higher<br />

salt concentrations, an excessive shielding <strong>and</strong> further enhancement<br />

<strong>of</strong> the solution hydrophobicity due to salt excess occurs,<br />

which induces a reduction <strong>of</strong> the ThT fluorescence by (a)<br />

decreasing the fraction <strong>of</strong> formed fibrils due to an unsuitable<br />

energy balance to establish effective intermolecular interactions,<br />

which favors r<strong>and</strong>om aggregation; (b) favoring side-to-side<br />

adsorption between protein molecules <strong>and</strong> fibrils, which leads<br />

to the formation fibril bundling, as observed by TEM; or (c)<br />

blocking the ThT binding sites on the fibrils due to a reduction<br />

in exposed surface area. Thus, there is a change in the energy<br />

l<strong>and</strong>scape <strong>of</strong> the fibrillation process. CD, FT-IR, <strong>and</strong> Trp<br />

fluorescence spectra seem to confirm this picture by monitoring<br />

the structural changes in both tertiary <strong>and</strong> secondary structure<br />

along the fibrillation process. In this way, a reduction in -sheet<br />

structure at large salt concentrations is also confirmed from the<br />

analysis <strong>of</strong> far-UV CD spectra. In contrast, under acidic<br />

conditions, a progressive enhancement <strong>of</strong> HSA fibrillation is<br />

observed as the electrolyte concentration in solution increases.<br />

Both the distinct ionization state <strong>and</strong> the different initial<br />

structural state <strong>of</strong> the protein at acidic conditions (exp<strong>and</strong>ed<br />

E-state) may be the origin <strong>of</strong> this behavior. On the other h<strong>and</strong>,<br />

we have noted that the fibrillation process <strong>of</strong> HSA does not<br />

show a lag-phase, except at acidic pH in the absence <strong>of</strong> added<br />

salt. This involves that the initial aggregation is a downhill<br />

process which does not require a highly organized <strong>and</strong> unstable<br />

nucleus, with a progressive increase <strong>of</strong> the -sheet (up to 26%)<br />

<strong>and</strong> an unordered conformation at the expense <strong>of</strong> the R-helical<br />

conformation. Finally, the obtained fibrils show a curly morphology<br />

<strong>and</strong> differ in length depending on solution conditions.<br />

In fact, at acidic conditions, a progressive evolution <strong>of</strong> the<br />

resulting fibrils can be found from a seam <strong>of</strong> -sheet structure<br />

decorated by relatively disorganized R-helical structure to highly<br />

structured fibrils, as found under physiological conditions, as<br />

derived from CD <strong>and</strong> FT-IR data.<br />

Acknowledgment. The authors thank the Ministerio de<br />

Ciencia e Innovación for support via Project MAT-2007-61604.<br />

The authors also thank Dr. Eugenio Vázquez for his assistance<br />

with CD measurements.<br />

Supporting Information Available: Figures showing a<br />

difference absorbance spectrum <strong>of</strong> Congo Red, time evolution<br />

172


Fibrillation Process <strong>of</strong> Human Serum Albumin J. Phys. Chem. B, Vol. 113, No. 30, 2009 10529<br />

<strong>of</strong> ThT fluorescence, <strong>and</strong> FT-IR spectra <strong>of</strong> native HSA. This<br />

material is available free <strong>of</strong> charge via the Internet at http://<br />

pubs.acs.org.<br />

References <strong>and</strong> Notes<br />

(1) Van der Linden, E.; Venema, P. Curr. Opin. Colloid Sci. 2007,<br />

12, 158.<br />

(2) Shahi, P.; Sharma, R.; Sanger, S.; Kumar, I.; Jolly, R. S.<br />

Biochemistry 2007, 46, 7365.<br />

(3) Dobson, C. M. Nature 2003, 426, 884.<br />

(4) Fändrich, M. Cell. Mol. Life Sci. 2007, 64, 2066.<br />

(5) Wang, W. Int. J. Pharm. 2005, 289, 1.<br />

(6) Gazit, E. FEBS J. 2007, 274, 317.<br />

(7) Chiti, F.; Webster, P.; Taddei, N.; Clark, A.; Stefani, M.; Ramponi,<br />

G.; Dobson, C. M. Proc. Natl. Acad. Sci. U.S.A. 1999, 96, 3590.<br />

(8) Cherny, I.; Gazit, E. Angew. Chem., Int. Ed. 2008, 48, 4062.<br />

(9) Gras, S. L.; Tickler, A. K.; Squieres, A. M.; Devlin, G. L.; Horton,<br />

M. A.; Dobson, C. M.; MacPhee, C. E. Biomaterials 2008, 29, 1553.<br />

(10) Dobson, C. M. Trends Biochem. Sci. 1999, 24, 329.<br />

(11) Thirumulai, D.; Klimov, D. K.; Dima, R. I. Curr. Opin. Struct.<br />

Biol. 2003, 13, 146.<br />

(12) Qu, Y.; Bolen, C. L.; Bolen, D. W. Proc. Natl. Acad. Sci. U.S.A.<br />

1998, 95, 9268.<br />

(13) Chiti, F.; Taddei, N.; Bucciantini, M.; White, P.; Ramponi, G.;<br />

Dobson, C. M. EMBO J. 2000, 19, 1441.<br />

(14) Niraula, T. N.; Haraoka, K.; Ando, Y.; Li, H.; Yamada, H.; Akasaka,<br />

K. J. Mol. Biol. 2002, 320, 333.<br />

(15) Krebs, M. R.; Wilkins, D. K.; Chung, E. W.; Pikeathly, M. C.;<br />

Chamberlain, A. K.; Zurdo, J.; Robinson, C. V.; Dobson, C. M. J. Mol.<br />

Biol. 2000, 300, 541.<br />

(16) Moraitakis, G.; Goodfellow, J. M. Biophys. J. 2003, 84, 2149.<br />

(17) De Felice, F. G.; Vieira, M. N.; Meirelles, M. N.; Morozova-Roche,<br />

L. A.; Dobson, C. M.; Ferreira, S. T. FASEB J. 2004, 18, 1099.<br />

(18) Ferrao-Gonzales, A. D.; Souto, S. O.; Silva, J. L.; Foguel, D. Proc.<br />

Natl. Acad. Sci. U.S.A. 2000, 97, 6445.<br />

(19) Khurana, R.; Gillespie, J. R.; Talapatra, A.; Minert, L. J.; Ionescu-<br />

Zanetti, C.; Mollet, I.; Fink, A. L. Biochemistry 2001, 40, 3525.<br />

(20) Zurdo, J.; Guijarro, J. I.; Jiménez, J. L.; Saibil, H. R.; Dobson,<br />

C. M. J. Mol. Biol. 2001, 311, 325.<br />

(21) Uversky, V. N.; Fink, A. L. Biochim. Biophys. Acta 2004, 1698,<br />

131.<br />

(22) Plakoutsi, G.; Taddei, N.; Stefani, M.; Chiti, F. J. Biol. Chem. 2004,<br />

279, 14111.<br />

(23) Eakin, C. M.; Attenello, F. J.; Morgan, C. J.; Miranker, A. D.<br />

Biochemistry 2004, 35, 6470.<br />

(24) Fändrich, M. C.; Dobson, C. M. EMBO J. 2002, 21, 5682.<br />

(25) Timasheff, S. N. Annu. ReV. Biophys. Biomol. Struct. 1993, 22,<br />

67.<br />

(26) Hoyer, W.; Anthony, T.; Cherny, D.; Heim, G.; Jovin, T. M.;<br />

Subramaniam, V. J. Mol. Biol. 2002, 322, 383.<br />

(27) Chiti, F.; Calamai, M.; Taddei, N.; Stefani, M.; Ramponi, G.;<br />

Dobson, C. M. Proc. Natl. Acad. Sci. 2002, 99, 16419.<br />

(28) Taboada, P.; Barbosa, S.; Castro, E.; Mosquera, V. J. Phys. Chem.<br />

B 2006, 110, 20733.<br />

(29) Taboada, P.; Barbosa, S.; Castro, E.; Gutiérrez-Pichel, M.; Mosquera,<br />

V. Chem. Phys. 2007, 340, 59.<br />

(30) Pace, C. N.; Vajdos, F.; Fee, L.; Grimsley, G.; Gray, T. Protein<br />

Sci. 1995, 4, 2411.<br />

(31) Sreerama, N.; Woody, R. W. Anal. Biochem. 2000, 287, 252.<br />

(32) Holm, N. K.; Jespersen, S. K.; Thomassen, L. V.; Wolff, T. Y.;<br />

Sehgal, P.; Thomsen, L. A.; Christiansen, G.; Andersen, C. B.; Knudsen,<br />

A. D.; Otzen, D. E. Biochim. Biophys. Acta 2007, 1774, 1128.<br />

(33) Pearce, F. G.; Mackintosh, S. H.; Gerrard, J. A. J. Agric. Food<br />

Chem. 2007, 55, 318.<br />

(34) Gorinstein, S.; Caspi, A.; Rosen, A.; Goshev, I.; Zemser, M.; Weisz,<br />

M.; Añon, M. C.; Libman, I.; Lerner, H. T.; Trakhtenberg, S. J. Peptide<br />

Res. 2002, 59, 71.<br />

(35) Juárez, J.; Taboada, P.; Mosquera, V. Biophys. J. 2009, 96, 2353.<br />

(36) Sagis, L. M. C.; Veerman, C.; Van der Linden, E. Langmuir 2004,<br />

20, 924.<br />

(37) Shaw, A. K.; Pal, S. K. J. Photochem. Photobiol. B 2008, 90, 69.<br />

(38) Muzammil, S.; Kumar, Y.; Tayyal, S. Eur. J. Biochem. 1999, 266,<br />

26.<br />

(39) Pereira, L. G. C.; Théodoly, O.; Blanch, H. W.; Radke, C. J.<br />

Langmuir 2003, 19, 2349.<br />

(40) Farruggia, B.; Rodriguez, F.; Rigatuso, R.; Fidelio, G.; Picó, G. J.<br />

Protein Chem. 2001, 20, 81.<br />

(41) Flora, K.; Brennan, J. D.; Baker, G. A.; Doody, M. A.; Brigth,<br />

F. V. Biophys. J. 1998, 75, 1084.<br />

(42) Tavirani, M. R.; Moghaddamnia, S. H.; Ranjbar, B.; Amani, M.;<br />

Marashi, S. A. J. Biochem. Mol. Biol. 2006, 39, 530.<br />

(43) Rader, A. J.; Hespenheide, B. M.; Kuhn, K. A.; Thorpe, M. F. Proc.<br />

Natl. Acad. Sci U.S.A. 2002, 99, 3540.<br />

(44) Ferrone, F. Methods Enzymol. 1999, 309, 256.<br />

(45) Lin, M.-S.; Chen, L.-Y.; Tsai, H.-T.; Wang, S. S.-S.; Chang, Y.;<br />

Higuchi, A.; Chen, W.-Y. Langmuir 2008, 24, 5802.<br />

(46) Hills, R. D.; Brooks, C. L. J. Mol. Biol. 2007, 368, 894.<br />

(47) Hong, Y.; Pritzker, M. D.; Legge, R. L.; Chen, P. Colloids Surf.,<br />

B 2005, 46, 152.<br />

(48) Fujiwara, S.; Matsumoto, F.; Yonezawa, Y. J. Mol. Biol. 2003,<br />

331, 21.<br />

(49) Sikkink, L. A.; Ramírez-Alvarado, M. Biophys. Chem. 2008, 135,<br />

25.<br />

(50) Zheng, J.; Jang, H.; Nussinov, R. Biochemistry 2008, 47, 2497.<br />

(51) Cerda-Costa, N.; Esteras-Chopo, A.; Aviles, F. X.; Serrano, L.;<br />

Villegas, V. J. Mol. Biol. 2007, 366, 1351.<br />

(52) Plakoutsi, G.; Bemporad, F.; Calamai, M.; Taddei, N.; Dobson,<br />

C. M.; Chiti, F. J. Mol. Biol. 2005, 351, 910.<br />

(53) Sreerama, N.; Woody, R. W. Protein Sci. 2003, 12, 384.<br />

(54) Kheterpal, I.; Chen, M.; Cook, K. D.; Wetzel, R. J. Mol. Biol. 2006,<br />

361, 785.<br />

(55) Myers, S. L.; Thomson, N. H.; Radford, S. E.; Ashcr<strong>of</strong>t, A. E. Rapid<br />

Commun. Mass Spectrom. 2006, 20, 1628.<br />

(56) Casal, H. L.; Kohler, U.; Mantsch, H. H. Biochim. Biophys. Acta<br />

1988, 957, 11.<br />

(57) Fabian, H.; Choo, L.-P.; Szendrei, G. I.; Jackson, M.; Halliday,<br />

W. C.; Otvos, L., Jr.; Mantsch, H. H. Appl. Spectrosc. 1993, 47, 1513.<br />

(58) Sunde, M.; Blake, C. AdV. Protein Chem. 1997, 50, 123.<br />

(59) Bouchard, M.; Zurdo, J.; Nettleton, E. J.; Dobson, C. M.; Robinson,<br />

C. V. Protein Sci. 2000, 9, 1960.<br />

(60) Sibley, S. P.; Sosinsky, K.; Gulian, L. E.; Gibbs, E. J.; Pasternack,<br />

R. F. Biochemistry 2008, 47, 2858.<br />

(61) Sambashivan, S.; Liu, M. R.; Sawaya, M. R.; Gingery, M.;<br />

Eisenberg, D. Nature 2005, 437, 266.<br />

(62) Ranson, N.; Stromer, T.; Bousset, L.; Melki, R.; Serpell, R. C.<br />

Protein Sci. 2006, 15, 2481.<br />

(63) Ramírez-Alvarado, M.; Merkel, J. S.; Regan, L. Proc. Natl. Acad.<br />

Sci. U.S.A. 2000, 97, 8979.<br />

(64) Damaschun, G.; Damaschun, H.; Fabian, H.; Gast, K.; Krober, R.;<br />

Wieske, M.; et al. Proteins: Struct., Funct., Genet. 2000, 39, 204.<br />

(65) Goda, S.; Takano, K.; Yamagata, Y.; Nagata, R.; Akutsu, H.; Maki,<br />

S.; et al. Protein Sci. 2000, 9, 369–375.<br />

JP902224D<br />

173


Additional Supra-<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> Human Serum Albumin under Amyloid-Like-Forming<br />

Solution Conditions<br />

Introduction<br />

Josué Juárez, † Pablo Taboada,* ,† Sonia Goy-López, † Adriana Cambón, †<br />

Marie-Beatrice Madec, ‡ Stephen G. Yeates, ‡ <strong>and</strong> Víctor Mosquera †<br />

Grupo de Física de Coloides y Polímeros, Departamento de Física de la Materia Condensada,<br />

Facultad de Física, UniVersidad de Santiago de Compostela, E-15782 Santiago de Compostela, Spain, <strong>and</strong><br />

Organic Materials InnoVation Center, School <strong>of</strong> Chemistry, UniVersity <strong>of</strong> Manchester, Manchester M13 9PL,<br />

United Kingdom<br />

ReceiVed: May 5, 2009; ReVised Manuscript ReceiVed: June 22, 2009<br />

Protein aggregation has a multitude <strong>of</strong> consequences ranging from affecting protein expression to its implication<br />

in different diseases. Of recent interest is the specific form <strong>of</strong> aggregation leading to the formation <strong>of</strong> amyloid<br />

fibrils, structures associated with diseases such as Alzheimer’s disease. These fibrils can further associate in<br />

other more complex structures such as fibrillar gels, plaques, or spherulitic structures. In the present work,<br />

we describe the physical <strong>and</strong> structural properties <strong>of</strong> additional supraself-assembled structures <strong>of</strong> human serum<br />

albumin under solution conditions in which amyloid-like fibrils are formed. We have detected the formation<br />

<strong>of</strong> ordered aggregates <strong>of</strong> amyloid fibrils, i.e., spherulites which possess a radial arrangement <strong>of</strong> the fibrils<br />

around a disorganized protein core <strong>and</strong> sizes <strong>of</strong> several micrometers by means <strong>of</strong> polarized optical microscopy,<br />

laser confocal microscopy, <strong>and</strong> transmission electron microscopy. These spherulites are detected both in solution<br />

<strong>and</strong> embedded in an isotropic matrix <strong>of</strong> fibrillar gels. In this regard, we have also noted the formation <strong>of</strong><br />

protein gels when the protein concentration <strong>and</strong>/or ionic strength exceds a threshold value (the gelation point)<br />

as observed by rheometry. Fibrillar gels are formed through intermolecular nonspecific association <strong>of</strong> amyloid<br />

fibrils at a pH far away from the isolectric point <strong>of</strong> the protein where protein molecules seem to display a<br />

“solid-like” behavior due to the existence <strong>of</strong> non-DLVO (Derjaguin-L<strong>and</strong>au-Verwey-Overbeck) intermolecular<br />

repulsive forces. As the solution ionic strength increases, a coarsening <strong>of</strong> this type <strong>of</strong> gel is observed<br />

by environmental scanning microscopy. In contrast, at pH close to the protein isoelectric point, particulate<br />

gels are formed due to a faster aggregation process, which does not allow substantial structural reorganization<br />

to enable the formation <strong>of</strong> ordered structures. This behavior also additionally corroborates that the existence<br />

<strong>of</strong> particulates might also be a generic property <strong>of</strong> all polypeptide chains as amyloid fibril formation under<br />

suitable conditions.<br />

A number <strong>of</strong> neurodegenerative diseases have been related<br />

to protein aggregation <strong>and</strong> deposition in the form <strong>of</strong> amyloid<br />

fibrils, such as type II diabetes, Alzehimer’s, Parkinson’s,<br />

Huntington’s, <strong>and</strong> prion diseases amongst others. 1-6 A wide<br />

range <strong>of</strong> proteins is known to misfold <strong>and</strong> aggregate in mildly<br />

or hard denaturating conditions as amyloid fibrils provided that<br />

the ability to fibrillate is independent <strong>of</strong> the original native<br />

structure <strong>of</strong> the protein, whose amino acid sequence primarily<br />

appears to play a key role in terms <strong>of</strong> filament arrangement, 7<br />

fibrillation kinetics, 8 <strong>and</strong> overall yield <strong>and</strong> stability <strong>of</strong> the<br />

fibrils. 9,10 These fibrils have been found to be based on a<br />

common structural motif consisting <strong>of</strong> continuous intermolecular<br />

-sheets which run along the fibril axis such that the individual<br />

-str<strong>and</strong>s are perpendicular to the fibril axis. 11<br />

Under certain conditions, protein fibrils can further associate<br />

either nonspecifically such as in fibrillar gels or plaques or in a<br />

more ordered fashion such as in spherulites. 12-18 It is known<br />

that fibrillar gels are formed by protein aggregation if the protein<br />

concentration exceeds a given critical value under conditions<br />

* Author to whom correspondence should be addressed. E-mail:<br />

pablo.taboada@usc.es.Phone:0034981563100,Ext.14042.Fax:0034981520676.<br />

† Universidad de Santiago de Compostela.<br />

‡ University <strong>of</strong> Manchester.<br />

J. Phys. Chem. B 2009, 113, 12391–12399 12391<br />

where protein molecules are partially denaturated (as at high<br />

temperatures) <strong>and</strong> electrically charged (this is, far away from<br />

their isolectric point) upon incubation, i.e., under conditions<br />

where fibrillation can occur. Nevertheless, the correspondence<br />

between fine str<strong>and</strong> gels <strong>and</strong> amyloid fibril structure was<br />

established recently. 19,20 In contrast, by decreasing the intermolecular<br />

repulsion through shifting the solution pH to values<br />

near the protein isoelectric point, or by increasing the solution<br />

ionic strength the gel networks are comprised <strong>of</strong> particulate<br />

aggregates.<br />

On the other h<strong>and</strong>, for systems where fibrillar protein aggregates<br />

<strong>and</strong> gels are formed, the str<strong>and</strong>s can further associate in a more<br />

ordered fashion such as spherulitic structures. These are characterized<br />

by the presence <strong>of</strong> anisotropic material ordered in a spherically<br />

symmetric way. This gives rise to the appearance <strong>of</strong> a Maltese<br />

cross extinction pattern when these structures are studied under a<br />

polarized light microscope, much in the same way as spherulites<br />

formed by synthetic polymers, 21,22 natural polymers, <strong>and</strong> biopolymers.<br />

23-25 In solutions in Vitro, spherulites have been reported for<br />

different synthetic <strong>and</strong> natural peptides, 17,26-30 <strong>and</strong> have also usually<br />

been associated to the amyloid-like formation pathways <strong>of</strong><br />

several proteins such as -lactoglobulin, 20,31,32 R-L-iduronidase, 33<br />

lysozyme, 18 <strong>and</strong> insulin. 16,34 Spherulites were also found in ViVo<br />

in Alzheimer’s disease, Gerstmann-Straulsler-Scheinker dis-<br />

10.1021/jp904167e CCC: $40.75 © 2009 American Chemical Society<br />

Published on Web 08/14/2009<br />

175


12392 J. Phys. Chem. B, Vol. 113, No. 36, 2009 Juárez et al.<br />

ease, Down’s syndrome, 35 <strong>and</strong> in tumors. 36,37 Thus, the formation<br />

<strong>of</strong> fibrillar structures may not be the end <strong>of</strong> the aggregation<br />

process, <strong>and</strong> proper knowledge <strong>of</strong> the nature <strong>and</strong> structure <strong>of</strong><br />

these supramolecular assemblies may have important implications<br />

in different fields such as medicine, 3,4,35-37 food industry, 38,39<br />

or biomaterials. 40<br />

In recent reports, 41,42 we have shown that partially destabilized<br />

human serum albumin (HSA) molecules form amyloid-like<br />

fibrils under different solution conditions. These fibrils feature<br />

the structural characteristics for amyloids: X-ray diffraction<br />

patterns, affinity to Congo Red <strong>and</strong> Thi<strong>of</strong>lavin T dyes, birefringence,<br />

<strong>and</strong> high stability. Nevertheless, different fibrillation<br />

pathways, fibrillation kinetics, <strong>and</strong> structural intermediates <strong>and</strong><br />

morphological differences in the resulting fibrils were observed<br />

depending on the incubation conditions. 42<br />

In this paper, we extend our previous studies in order to<br />

explore the solution conditions which enable the formation <strong>of</strong><br />

different supramolecular assemblies <strong>of</strong> HSA amyloid-like fibers,<br />

<strong>and</strong> the structure <strong>of</strong> these resulting assemblies. To do that,<br />

extensive incubation <strong>of</strong> HSA solutions <strong>of</strong> pH 2.5, 5.5, <strong>and</strong> 7.4<br />

at elevated temperature <strong>and</strong> ionic strengths ranging from 0 to<br />

250 mM NaCl was made, <strong>and</strong> the resulting protein assemblies<br />

were analyzed by optical, confocal, transmission electron,<br />

environmental scanning electron, <strong>and</strong> atomic force microscopies,<br />

rheometry, <strong>and</strong> attenuated reflectance Fourier transform infrared<br />

(ATR-FTIR) <strong>and</strong> fluorescence spectroscopies. We have observed<br />

that spherulitic formation takes place under conditions where<br />

amyloid fibrils were previously present. The spherulites possess<br />

a radial arrangement <strong>of</strong> the protein fibrils around a disorganized<br />

protein core <strong>and</strong> sizes <strong>of</strong> 5-50 µm. In addition, different types<br />

<strong>of</strong> gels were also obtained upon extensive incubation under<br />

different solution conditions, either with fibrillar or particulate<br />

structure. Fibrillar gels that form at a pH far from the isoeletric<br />

point <strong>of</strong> the protein display a “solid-like” behavior <strong>and</strong> are<br />

formed through intermolecular nonspecific association <strong>of</strong> amyloid<br />

fibrils. A coarsening <strong>of</strong> these gels occurs when the solution<br />

ionic strength increases. When the solution pH is close to the<br />

protein isoelectric point, particulate gels are formed. These<br />

originate from a faster protein aggregation which does not allow<br />

the necessary structural reorganization to enable the formation<br />

<strong>of</strong> more ordered structures.<br />

Materials <strong>and</strong> Methods<br />

Materials. Human serum albumin (70024-90-7) <strong>and</strong> Thi<strong>of</strong>lavin<br />

T (ThT) were obtained from Sigma Chemical Co. <strong>and</strong> used<br />

as received. All other chemicals were <strong>of</strong> the highest purity<br />

available.<br />

Preparation <strong>of</strong> HSA Solutions. Protein was used after further<br />

purification by liquid chromatography using a Superdex 75<br />

column equilibrated with 0.01 M phosphate. Experiments were<br />

carried out using double distilled, deionized, <strong>and</strong> degassed water.<br />

The buffer solutions used were glycine + HCl (I ) 0.01 M)<br />

for pH 2.5, sodium acetate-acetic acid for pH 5.5 (I ) 0.01<br />

M), <strong>and</strong> sodium monophosphate-sodium diphosphate for pH<br />

7.4 (I ) 0.01 M), respectively. HSA was dissolved in each buffer<br />

solution to a final concentration <strong>of</strong> typically 20 mg/mL if not<br />

otherwise stated, <strong>and</strong> dialyzed extensively against buffer solution.<br />

Protein concentration was determined spectrophotometrically,<br />

using a molar absorption coefficient <strong>of</strong> 35219 M-1 cm-1 at 280 nm. 43 Prior to incubation, solutions were filtered through<br />

a 0.2 µm filter into sterile test tubes. Samples were incubated<br />

at a specified temperature (65 °C) for up to 15 days in a refluxed<br />

reactor if not otherwise stated. Samples were taken out at<br />

intervals <strong>and</strong> stored on ice before adding ThT.<br />

Polarized Optical Microscopy (POM). After incubation,<br />

aliquots <strong>of</strong> protein solutions were removed from the vials <strong>and</strong><br />

put onto microscope slides. These were immediately studied<br />

using a Zeiss Axioplan optical microscope (Carl Zeiss Ltd.,<br />

Welwyn Garden City, U.K.). A total magnification <strong>of</strong> either<br />

50× or 100× was used. A polarizer <strong>and</strong> analyzer were put in<br />

fixed positions, orthogonal to one another for the cross-polarizer<br />

imaging. Images were taken using a Kodak digitial camera<br />

mounted on the top <strong>of</strong> the microscope. The size <strong>of</strong> the<br />

spherulites seen in the images was determined by calibrating<br />

with a scale bar <strong>of</strong> known dimensions <strong>and</strong> thereby quantifying<br />

the spatial resolution per pixel image.<br />

Laser Confocal Microscopy. Confocal microscopy was<br />

performed on a Leica TCS SP2 confocal system mounted on a<br />

Leica DM-IRE2 upright microscope, using a 40× objective. For<br />

fluorescence measurements, Thi<strong>of</strong>lavin T was added to spherulitic<br />

solutions <strong>and</strong> excited using the 458 nm line <strong>of</strong> an argon<br />

ion laser. Simultaneously, transmission images were obtained<br />

with crossed polarizers in place using the 633 nm line <strong>of</strong> a<br />

He-Ne laser.<br />

Thi<strong>of</strong>lavin T Spectroscopy. Protein <strong>and</strong> ThT were dissolved<br />

in proper buffer at a final protein-dye molar ratio <strong>of</strong> 50:1.<br />

Samples were continuously stirred during measurements. Fluorescence<br />

was measured in a Cary Eclipse fluorescence spectrophotometer<br />

equipped with a temperature control device <strong>and</strong><br />

a multicell sample holder (Varian Instruments Inc.). Excitation<br />

<strong>and</strong> emission wavelengths were 450 <strong>and</strong> 482 nm, respectively.<br />

All intensities were background-corrected for the ThT fluorescence<br />

in the respective solvent without the protein. Data from<br />

at least three different separate experiments were averaged <strong>and</strong><br />

normalized on a scale from 0 to 1.<br />

Attenuated Total Reflectance Fourier Transform Infrared<br />

Spectroscopy (ATR-FTIR). ATR-FTIR spectra <strong>of</strong> HSA in<br />

aqueous solutions were determined by using a FTIR spectrometer<br />

(model IFS-66v from Bruker) equipped with a horizontal<br />

ZnS ATR accessory. The spectra were obtained at a resolution<br />

<strong>of</strong>2cm -1 , <strong>and</strong> generally 200 scans were accumulated to get a<br />

reasonable signal-to-noise ratio. Solvent spectra were also<br />

examined using the same accessory <strong>and</strong> instrument conditions.<br />

Each different sample spectrum was obtained by digitally<br />

subtracting the solvent spectrum from the corresponding sample<br />

spectrum. Each sample solution was repeated three times to<br />

ensure reproducibility <strong>and</strong> averaged to produce a single spectrum.<br />

Transmission Electron Microscopy (TEM). Suspensions <strong>of</strong><br />

HSA were applied to carbon-coated copper grids, blotted,<br />

washed, negatively stained with 2% (w/v) phosphotungstic acid,<br />

air-dried, <strong>and</strong> then examined with a Phillips CM-12 transmission<br />

electron microscope operating at an accelerating voltage <strong>of</strong> 120<br />

kV. Samples were diluted between 20-200-fold <strong>and</strong> sonicated<br />

when needed prior to deposition on the grids.<br />

Environmental Scanning Electron Microscopy (ESEM).<br />

Suspensions <strong>of</strong> HSA were applied to glass-coated stainless steel<br />

grids, blotted, washed, air-dried, <strong>and</strong> then examined with an<br />

LEO-435 VP scanning electron microscope (Leica Mycrosystems,<br />

Cambridge, U.K.) operating at an accelerating voltage <strong>of</strong><br />

20 kV. Samples were diluted when needed prior to deposition<br />

on the grids. Samples were left to equilibrate at 2 °C. A few<br />

drops <strong>of</strong> distilled <strong>and</strong> deionized water were place around the<br />

sample. The chamber was flooded repeatedly with water vapor<br />

<strong>and</strong> the pressure subsequently reduced to ∼5 Torr. Further<br />

decreases in pressure led to evaporation <strong>of</strong> water <strong>and</strong> drying <strong>of</strong><br />

the samples.<br />

Atomic Force Microscopy (AFM). AFM images were<br />

recorded in the tapping mode by using an AFM XE100<br />

176


Additional Supra-<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> HSA J. Phys. Chem. B, Vol. 113, No. 36, 2009 12393<br />

instrument (PSIA Inc., Seoul, Korea) equipped with a rigid<br />

silicon cantilever <strong>of</strong> nominal spring constant <strong>of</strong> about 40 mN/m<br />

<strong>and</strong> resonant frequency <strong>of</strong> 325 kHz. Immediately after incubation,<br />

protein samples were diluted 20-400 times onto freshly<br />

cleaved muscovite mica (from Sigma Chemical Co.) attached<br />

to a magnetic steel disk which served as a sample holder. The<br />

abrupt dilution <strong>of</strong> the samples immediately quenches the<br />

concentration-dependent aggregation process. The AFM samples<br />

were dried in air or under nitrogen flow when required. Control<br />

samples (freshly cleaved mica, mica <strong>and</strong> buffer solution) were<br />

also investigated with AFM to exclude possible artifacts.<br />

Topography <strong>and</strong> phase-shift data were collected in the trace <strong>and</strong><br />

retrace direction <strong>of</strong> the raster, respectively. The <strong>of</strong>fset point was<br />

adapted accordingly to the roughness <strong>of</strong> the sample. The scan<br />

rate was tuned proportionally to the area scanned <strong>and</strong> kept within<br />

0.35-2 Hz range.<br />

Rheometry. The rheological properties <strong>of</strong> the samples were<br />

determined using a Bohlin CS10 rheometer with water bath<br />

temperature control. Couette geometry (bob, 24.5 mm in<br />

diameter, 27 mm in height; cup, 26.5 mm in diameter, 29 mm<br />

in height) was used, with a 2.5 cm 3 sample being added to the<br />

cup in the mobile state. Samples <strong>of</strong> very high modulus were<br />

investigated using cone-<strong>and</strong>-plate geometry (diameter 40 mm,<br />

angle 4°). A solvent trap maintained a water-saturated atmosphere<br />

around the cells. Frequency scans <strong>of</strong> storage (G′) <strong>and</strong><br />

loss (G′′) moduli were recorded for selected protein concentrations<br />

<strong>and</strong> temperatures with the instrument in the oscillatory<br />

shear mode <strong>and</strong> with the strain amplitude (A) maintained at a<br />

low value (A < 0.5%) by means <strong>of</strong> the autostress facility <strong>of</strong> the<br />

Bohlin s<strong>of</strong>tware. This ensured that measurements <strong>of</strong> G′ <strong>and</strong> G′′<br />

were in the linear viscoelastic region. Measurements on solutions<br />

<strong>of</strong> low modulus (G′ ) 1-10 Pa) which fell outside the range<br />

for satisfactory autostress feedback were rejected.<br />

Results <strong>and</strong> Discussion<br />

Human serum albumin consists <strong>of</strong> 585 amino acids in a single<br />

polypeptide chain, with a globular structure composed <strong>of</strong> three<br />

main domains that are loosely joined together through physical<br />

forces <strong>and</strong> six subdomains that are wrapped by disulfide bonds.<br />

The protein contains 17 disulfide bridges <strong>and</strong> one free SH group.<br />

Most <strong>of</strong> the HSA sequence (>60%) is arranged in R-helix<br />

structure, with the subsequent tightening <strong>of</strong> its structure through<br />

intramolecular interactions such as hydrogen bonds. Nevertheless,<br />

amyloid-like formation <strong>of</strong> HSA has been reported under<br />

different solution conditions. 41,42 HSA fibril formation usually<br />

takes place in the absence <strong>of</strong> a lag phase, <strong>and</strong> the resulting fibrils<br />

share the common structural features <strong>of</strong> “true” amyloid fibrils,<br />

although their final morphology depends on the incubation<br />

conditions. 42<br />

Spherulite Formation <strong>and</strong> Characterization. The propensity<br />

for spherulite formation was investigated by using optical<br />

<strong>and</strong> confocal microscopy at pH 2.5, 5.5, <strong>and</strong> 7.4 at 65 °C upon<br />

incubation in the presence <strong>of</strong> NaCl in the range 0-250 mM.<br />

We have observed that spherulitic structures are found under<br />

conditions where HSA fibrils were previously detected, 42 that<br />

is, at physiological pH <strong>and</strong> under acidic conditions in the<br />

presence <strong>of</strong> the highest added salt concentrations (100-250 mM<br />

NaCl). Although data were not collected quantitatively, the<br />

highest density <strong>of</strong> spherulites was observed at pH 7.4 in the<br />

presence <strong>of</strong> 50 mM NaCl, <strong>and</strong> at acidic pH in the presence <strong>of</strong><br />

250 mM. This agrees with the maximum efficiency <strong>of</strong> fibril<br />

formation previously reported. 42 In this regard, it has been<br />

previously found41 that at both physiological pH <strong>and</strong> temperature<br />

fibril formation is disfavored at high ionic strengths due to an<br />

excesive shielding <strong>of</strong> repulsive electrostatic forces, which<br />

involves an increase in rapid r<strong>and</strong>om protein aggregation. In<br />

contrast, under acidic conditions fibril formation is continuosly<br />

enhanced as the added salt concentration increases. This<br />

difference has been assigned to the different nature <strong>and</strong> structure<br />

<strong>of</strong> the protein precursor states, which leads to different fibrillation<br />

rates, fibril conversion extents, fibrillation pathways, <strong>and</strong><br />

structures <strong>of</strong> resulting fibers. 44-46 In addition, no differences were<br />

observed between the spherulites obtained in the various samples<br />

studied regardless <strong>of</strong> the incubation conditions.<br />

Analysis <strong>of</strong> the solutions at both pH 2.5 <strong>and</strong> 7.4 by polarized<br />

optical microscopy allowed us to observe the presence <strong>of</strong><br />

classical Maltese cross extinction patterns (Figure 1a) as a key<br />

mark <strong>of</strong> the presence <strong>of</strong> spherulites. This type <strong>of</strong> pattern was<br />

not observed at pH 5.5. The observed spherulites possess sizes<br />

ranging from 5 to 50 µm. In addition, the test tubes when placed<br />

between crossed polarizers showed birefrigence from spherulites<br />

in the samples (not shown). Spherulitic structures are detected<br />

both in solution prior to gelification <strong>and</strong> embedded in an<br />

isotropic gel matrix after gelification. It should also be noted<br />

that nonspherulitic areas <strong>of</strong> the sample, which may contain fibrils<br />

not assembled into spherulites, are not visibly birefringent,<br />

indicating that polypeptide chains in this region are unoriented.<br />

Similar spherulitic occurrences in protein gels were also found,<br />

for example, for whey proteins, 47 -lactoglobulin, 29,30 <strong>and</strong><br />

insulin. 32,33<br />

To confirm that the spherulites contain amyloid fibrils, they<br />

were imaged by confocal microscopy using ThT as the<br />

fluorescent dye, whose fluorescence is known to increase<br />

markedly upong binding to amyloid fibrils. 48 In confocal mode,<br />

the fluorescence emission was observed to come from the<br />

spherulites (Figure 1b), which confirms they are composed <strong>of</strong><br />

amyloid fibers. 49 Confocal microscopy in transmission mode<br />

resulted in an image similar to that obtained with an optical<br />

microscope (Figure 1c), clearly displaying the presence <strong>of</strong><br />

spherulites as Maltese cross patterns <strong>of</strong> comparable sizes to those<br />

observed by optical miscroscopy. The nonordered <strong>and</strong> nonfluorescent<br />

core is probably due to nonpenetration <strong>of</strong> ThT. As this<br />

core is present even after extensive incubation in the presence<br />

<strong>of</strong> the dye, a more feasible explanation is the absence <strong>of</strong><br />

significant amounts <strong>of</strong> amyloid material in this core, that is, it<br />

is formed before spherulitic growth takes place due to an<br />

interplay between amyloid fibrillation <strong>and</strong> protein r<strong>and</strong>om<br />

aggregation. 16 Aliquots <strong>of</strong> spherulitic solutions were also added<br />

to solutions containing ThT. The enhanced fluorescence relative<br />

to solutions <strong>of</strong> ThT <strong>and</strong> ThT containing freshly dissolved HSA<br />

additionally proved the presence <strong>of</strong> amyloid structures forming<br />

the spherulites (see Figure 1d).<br />

To reveal the orientation <strong>of</strong> the fibrils within the spherulites,<br />

insertion <strong>of</strong> a waveplate at 45° angle to the cross polars provided<br />

confirmation <strong>of</strong> this radial arrangement (see inset in Figure 1a).<br />

The four quadrants in the spherulites are now colored. The<br />

appearance <strong>of</strong> the blue color indicates that the fast optical axis<br />

<strong>of</strong> the spherulites is aligned with the fast optical axis <strong>of</strong> the<br />

wave retardation plate. 34 The slow direction within the spherulite<br />

is the radial direction <strong>and</strong>, therefore, indicates that the fibrils<br />

lie radially within the spherulites, as seen previously for other<br />

proteins such as -lactoglobulin, 20,31,32 R-L-iduronidase, 33<br />

lysozyme, 18 <strong>and</strong> insulin. 16,34 Figure 1e shows that out from the<br />

core the radially oriented internal structure appears to be str<strong>and</strong>s<br />

<strong>of</strong> fibril bundles (because individual fibrils are too narrow to<br />

be seen optically) with slight curvature to their orientation <strong>and</strong><br />

without obvious branching. This is particularly viewable at the<br />

177


12394 J. Phys. Chem. B, Vol. 113, No. 36, 2009 Juárez et al.<br />

Figure 1. (a) Optical microscopy images <strong>of</strong> grown spherulites <strong>of</strong> HSA at pH 7.4 in the absence <strong>of</strong> added NaCl at 65 °C after 72 h <strong>of</strong> incubation<br />

under cross-polarizers. The inset indicates the radial arrangement <strong>of</strong> fibrils through insertion <strong>of</strong> a waveplate at 45° between the polarizers. (b)<br />

Fluorescence image in the confocal mode <strong>of</strong> spherulites using ThT as a probe. (c) Confocal image in the transmission mode <strong>of</strong> spherulites between<br />

cross-polarizers. (d) ThT fluorescence upon incubation in a solution without HSA, containing native HSA <strong>and</strong> containing spherulites. (e) Optical<br />

microscopy image <strong>of</strong> HSA spherulites without polarizers. Scale bars are (a) 50, (b) 200, <strong>and</strong> (c <strong>and</strong> e) 100 µm.<br />

outer edges <strong>of</strong> the spherulites, where the fibril bundles are<br />

slightly separated from one another.<br />

The spherical morphology <strong>of</strong> spherulites was also observed<br />

by ESEM (see Figure 2), with sizes similar to those derived<br />

from optical <strong>and</strong> confocal microscopy. Their surface appears<br />

to have a texture on the length scale <strong>of</strong> several micrometers<br />

but shows no obvious orientational order. Some cracks on the<br />

particle surface appeared as a consequence <strong>of</strong> the dehydration<br />

process in ESEM. Unfortunately, we could not observe if the<br />

spherulites formed display a radial internal order, since internal<br />

cracking <strong>of</strong> spherulitic structures was not achieved.<br />

Direct measurement <strong>of</strong> spherulitic solutions was not posible<br />

by TEM provided that a film was formed on the grids which<br />

do not allow visualization. In this way, dilution <strong>and</strong> sonication<br />

<strong>of</strong> the spherulitic solutions were performed in order to investigate<br />

the samples by TEM. This treatment resulted in the<br />

disruption <strong>of</strong> the spherulites <strong>and</strong> the appearance <strong>of</strong> amyloidlike<br />

fibril structures, as seen in Figure 2d. This fact further<br />

confirms that spherulites are mainly composed <strong>of</strong> amyloid fibrils.<br />

Gel Formation <strong>and</strong> Characterization. Prolonged incubation<br />

<strong>of</strong> solutions containing 0.3 mM HSA (or 0.8 mM where<br />

corresponds) at either pH 7.4, 5.5 or 2.5 at 65 °C in the presence<br />

<strong>of</strong> different amounts <strong>of</strong> added salt leads to the formation <strong>of</strong><br />

gels. In this work, we have only analyzed solution conditions<br />

previously studied for amyloid fibril formation. 41,42 Significant<br />

differences were observed in the association properties <strong>of</strong> the<br />

different protein solutions. In this way, we have observed that<br />

at pH 7.4 <strong>and</strong> 5.5 gels were formed at electrolyte concentrations<br />

larger than 20 mM with gelation times decreasing from several<br />

days to hours as the solution ionic strength increases (or the<br />

protein concentration rises). On the other h<strong>and</strong>, under acidic<br />

conditions gels were only formed at the highest electrolyte<br />

concentration analyzed, 250 mM NaCl at 0.3 mM HSA. In<br />

general, it seems that gel formation is favored by the screening<br />

<strong>of</strong> electrostatic interactions between protein molecules, which<br />

favors protein aggregation. 45,46 Nevertheless, protein interactions<br />

at acidic pH are weaker than those at physiological pH due to<br />

the initial different conformational state <strong>of</strong> the protein (Exp<strong>and</strong>ed-E<br />

state under acidic conditions), which imparts a larger<br />

solubility to the acid-denaturated protein molecules as a result<br />

<strong>of</strong> electrostatic repulsion between them due to an increase <strong>of</strong><br />

surface-charged groups.<br />

Figure 3a shows cure curves (G′, the storage modulus, versus<br />

time) for gelling HSA solutions at pH 7.4, 5.5, <strong>and</strong> 2.5 in the<br />

presence <strong>of</strong> 100 mM NaCl at 65 °C. To follow the protein<br />

gelation in a suitable time scale for the experiment, we increased<br />

the protein concentration up to 0.8 mM. The obtained cure data<br />

are typical <strong>of</strong> a sol-gel transition <strong>of</strong> biopolymer solutions: The<br />

elastic component <strong>of</strong> the shear modulus starts small but<br />

undergoes a sudden increase at a critical time <strong>and</strong>, then,<br />

ultimately almost levels <strong>of</strong>f at longer times. For pH 7.4 <strong>and</strong><br />

2.5, the behavior is similar to that expected for a strongly gelling<br />

178


Additional Supra-<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> HSA J. Phys. Chem. B, Vol. 113, No. 36, 2009 12395<br />

Figure 2. (a-c) ESEM images <strong>of</strong> different HSA spherulites incubated at pH 7.4 at 65 °C in the absence <strong>of</strong> added electrolyte along dehydration.<br />

Progressive dehydration allows a better visualization <strong>of</strong> the spherulites. In part c, some surface cracking can be observed. The scale bars are 200,<br />

100, <strong>and</strong> 50 µm in parts a, b, <strong>and</strong> c, respectively. (d) TEM image <strong>of</strong> HSA fibrils obtained after disruption <strong>of</strong> grown spherulites by extensive<br />

sonication.<br />

Figure 3. (a) G′ versus time (cure data) for a 0.8 mM HSA solution<br />

heated at 65 °C atpH(9) 7.4, (b) 2.5, <strong>and</strong> (4) 5.5 in the presence <strong>of</strong><br />

100 mM NaCl. (b) Time evolution <strong>of</strong> storage (G′, O) <strong>and</strong> loss (G′′, 9)<br />

moduli <strong>of</strong> a 0.8 mM HSA solution heated at 65 °C <strong>and</strong> pH 7.4 in the<br />

presence <strong>of</strong> 100 mM NaCl.<br />

biopolymer solution, with the divergence between G′ <strong>and</strong> G′′<br />

becoming larger with time <strong>and</strong> G′′ < G′. Another feature <strong>of</strong><br />

these cure curves is the “dip-down” in the modulus at long times,<br />

which may be indicative <strong>of</strong> slippage <strong>of</strong> the gel samples. 13<br />

Interestingly, G′, though small initially at both pHs, was larger<br />

than G′′ (the loss modulus) even before the gelling point (see<br />

Figure 3b as an example). This “solid-like” behavior <strong>of</strong> the<br />

starting protein solutions has been ascribed to colloid-like<br />

Figure 4. Frequency depedence <strong>of</strong> the storage (G′, closed symbols)<br />

<strong>and</strong> loss (G′′, open symbols) moduli <strong>of</strong> a 0.8 mM HSA solution at pH<br />

(b) 7.4, (9) 2.5, <strong>and</strong> (2) 5.5 in the presence <strong>of</strong> 100 mM NaCl at 65<br />

°C.<br />

structuring in the pregel solution due to nonclassical DLVO<br />

forces. 50,51 This is indicative <strong>of</strong> a different self-assembly<br />

mechanism <strong>and</strong>/or kinetics if compared to what was observed<br />

at pH 5.5, for which G′′ > G′ at short incubation times. At the<br />

latter pH, faster aggregation takes place since enhanced protein/<br />

protein interactions are favored due to the proximity <strong>of</strong> the<br />

solution pH to the protein isoelectric point. It has been pointed<br />

out that rapid formation <strong>of</strong> an appreciable amount <strong>of</strong> intermolecular<br />

-sheets favored by protein/protein interactions 52 involves<br />

very compact structures which precipitate, thereby limiting the<br />

amount <strong>of</strong> protein available to form protein/water networks<br />

necessary for robust gel formation 53 as in the present case. This<br />

fact is also corroborated comparing the lower G′ values at pH<br />

5.5 to those obtained at pH 7.4 <strong>and</strong> 2.5.<br />

Figure 4 shows the corresponding frequency spectrum <strong>of</strong> G′<br />

<strong>and</strong> G′′ for the gels after completion <strong>of</strong> the cure experiment.<br />

Samples at pH 2.5 <strong>and</strong> 7.4 show typical gel spectra. G′ is<br />

179


12396 J. Phys. Chem. B, Vol. 113, No. 36, 2009 Juárez et al.<br />

Figure 5. (a <strong>and</strong> b) ESEM <strong>and</strong> (c) AFM images showing progressive coarsening <strong>of</strong> fibril str<strong>and</strong>s <strong>of</strong> HSA gels after incubation at 65 °C atpH7.4<br />

in the presence <strong>of</strong> 20, 100, <strong>and</strong> 250 mM NaCl, respectively (scale bars in parts a <strong>and</strong> b are 10 <strong>and</strong> 5 µm, respectively; AFM image size is 5 µm<br />

× 5 µm). (d) TEM <strong>and</strong> (e) AFM images <strong>of</strong> fibrils from coarsed HSA gels after disruption by dilution <strong>and</strong> sonication at physiological pH (AFM<br />

image size is 5 µm × 5 µm). (f <strong>and</strong> h) AFM <strong>and</strong> (g) ESEM images <strong>of</strong> particulate gels <strong>of</strong> HSA at pH 5.5 after incubation at 65 °C in the presence<br />

<strong>of</strong> 20 (f <strong>and</strong> g) <strong>and</strong> 150 mM NaCl (h) (scale bar in part g is 2 µm; AFM image size is 5 µm × 5 µm).<br />

essentially frequency-independent, so we can assert that the<br />

storage moduli will have a finite value at very long times (low<br />

frequencies); i.e., it confirms the solid-like properties consistent<br />

with the development <strong>of</strong> a network, as previously commented.<br />

On the other h<strong>and</strong>, G′ values at physiological pH are larger than<br />

those at acidic pH, which denotes the formation <strong>of</strong> a stiffer gel.<br />

This is also supported by the essentially frequency-independent<br />

behavior <strong>of</strong> G′′ at pH 7.4. In contrast, under acidic conditions,<br />

a slight minimum in G′′ can be observed, which becomes much<br />

more intense at pH 5.5. This minimum at pH 5.5 can be assigned<br />

to molecular packing, 54 <strong>and</strong> is in agreement with the lower G′<br />

values, the frequency-dependent behavior, <strong>and</strong> the particulate<br />

structure <strong>of</strong> the gels formed at this pH, as seen below.<br />

Figure 5 shows ESEM <strong>and</strong> AFM images <strong>of</strong> the gels formed<br />

at pH 7.4, 5.5, <strong>and</strong> 2.5 at different ionic strengths. For<br />

physiological conditions after incubation at 65 °C in the presence<br />

<strong>of</strong> 20 mM NaCl, a fibrillar gel formed by fine str<strong>and</strong>s can be<br />

observed (Figure 5a). These str<strong>and</strong>s are formed by interconnected<br />

amyloid-like fibrils as denoted by the cross- pattern<br />

derived from XRD analysis (see Figure S1 in the Supporting<br />

Information), <strong>and</strong> in agreement with the rheology data. Moreover,<br />

upon strong dilution <strong>and</strong> sonication the str<strong>and</strong>s dissociate<br />

into fibrils which showed a predominant curly morphology, as<br />

observed previously for HSA 42 <strong>and</strong> other proteins such as<br />

2 microglobulin, 55 R-crystallin, 56 or bovine serum albumin<br />

(BSA). 57 Nevertheless, some mature straight fibrils <strong>and</strong> ribbonlike<br />

structures could also be observed (see Figure S2 in the<br />

Supporting Information).<br />

When the ionic strength is increased at pH 7.4 up to 250<br />

mM NaCl, a coarsening <strong>of</strong> the gel is progressively observed,<br />

180


Additional Supra-<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> HSA J. Phys. Chem. B, Vol. 113, No. 36, 2009 12397<br />

with the presence <strong>of</strong> relatively long, large, <strong>and</strong> thick aggregates<br />

which seem to be composed, at least partially, <strong>of</strong> globular<br />

aggregates (Figure 5b <strong>and</strong> c). Additionally, the size <strong>of</strong> the<br />

elementary particles appeared to be larger than those formed at<br />

lower added salt concentrations. A similar picture was observed<br />

for the gel formed at acidic pH (not shown). Therefore, a<br />

transition from fine-str<strong>and</strong>ed to coarse gels with increasing ionic<br />

strength seems to be caused mainly by kinetic effects without<br />

accompanying fundamental changes in aggregation mechanisms.<br />

58 After extensive dilution <strong>and</strong> sonication, rupture <strong>of</strong> these<br />

types <strong>of</strong> aggregates in a variety <strong>of</strong> fibrillar structures rather than<br />

in smaller spherical aggregates was observed by TEM <strong>and</strong> AFM<br />

(see Figure 5d <strong>and</strong> e), as seen in previous studies. 59 The length<br />

<strong>of</strong> the fibrils varies between 0.1 <strong>and</strong> 1-2 µm with a thickness<br />

<strong>of</strong> 15-30 nm.<br />

In contrast, at pH 5.5 a network <strong>of</strong> quasipherical protein<br />

aggregates surrounded by liquid is formed; i.e., we observed a<br />

particulate gel (see Figure 5f-h). The particle size varies<br />

between ∼300 <strong>and</strong> ∼500 nm, which increases as the ionic<br />

strength increases. This leads to fused long <strong>and</strong> thick aggregates<br />

possibily composed <strong>of</strong> several particulates. In this regard, it has<br />

been speculated that intermolecular disulfide bonding is involved<br />

in connecting protein molecules within the particulates <strong>and</strong> that<br />

the connections among them is a nonspecific physical crosslinking<br />

without any specific connective sites. 58,60 ESEM also<br />

indicates that an amount <strong>of</strong> residual monomeric protein was<br />

left in the solutions (Figure 5g) <strong>and</strong> deposited between the<br />

particulates. This gives the appearance <strong>of</strong> connected particulates<br />

<strong>and</strong>, sometimes, apparently amorphous films. As commented<br />

above, shifting the pH toward the isoelectric point decreases<br />

the intermolecular electrostatic repulsion. This implies that<br />

aggregation becomes faster, preventing the formation <strong>of</strong> highly<br />

ordered nanostructures such as amyloid-like fibrils, but globular<br />

shapes are formed.<br />

Thus, it seems that the gelation <strong>of</strong> HSA is also pH-dependent<br />

<strong>and</strong> the structures which form the gel depend on the solution<br />

conditions. In this regard, in a recent report the capability <strong>of</strong><br />

several proteins such as -lactoglobulin, BSA, insulin, <strong>and</strong><br />

lysozyme among others to form either fibrillar or particulate<br />

gels under partially denaturing conditions but at different pHs<br />

has been confirmed. 61 This suggests that the formation <strong>of</strong><br />

particulates can be a generic property <strong>of</strong> all polypeptide chains<br />

like amyloid fibrillation is. In this way, the results reported in<br />

this work support this view.<br />

Finally, in order to investigate the internal structure <strong>of</strong> the<br />

fibrillar <strong>and</strong> particulate gels, attenuated total reflectance Fourier<br />

transform infrared spectroscopy (ATR-FTIR) measurements<br />

were made. All spectra were well resolved with clearly<br />

distinguishable secondary structure signatures. In this way,<br />

before incubation two major b<strong>and</strong>s peaks in the second<br />

derivative IR spectra in the spectral region <strong>of</strong> interest were<br />

observed at pH 7.4: the amide I b<strong>and</strong> at 1652 cm -1 <strong>and</strong> the<br />

amide II b<strong>and</strong> at 1544 cm -1 . This indicates the predominant<br />

structural contribution <strong>of</strong> major R-helix <strong>and</strong> minor r<strong>and</strong>om coil<br />

structures. 62,63 For the amide I b<strong>and</strong> (see Figure 6a), a shoulder<br />

at ca. 1630 cm -1 can also be observed in the second derivative<br />

spectra, which is related to a low intramolecular -sheet content.<br />

Additional peaks at ca. 1689 <strong>and</strong> 1514 cm -1 would correspond<br />

to -turn <strong>and</strong> tyrosine absorption, respectively. 63 As the pH is<br />

lowered, the remaining presence <strong>of</strong> the amide I <strong>and</strong> II b<strong>and</strong>s<br />

confirms that there is still a significant amount <strong>of</strong> R-helices,<br />

even at the most acidic pH (see Figure 6a).<br />

Following incubation at 65 °C <strong>and</strong> formation <strong>of</strong> the gel at<br />

pH 7.4, a red-shift <strong>of</strong> the amide I b<strong>and</strong> from 1652 to 1658 cm -1<br />

Figure 6. Second derivative <strong>of</strong> FTIR spectra <strong>of</strong> HSA samples formed<br />

at (a) pH 2.5, (b) pH 5.5, <strong>and</strong> (c) pH 7.4 in the presence <strong>of</strong> 50 mM<br />

NaCl (s) before <strong>and</strong> ( ···) after incubation <strong>and</strong> gelation at 65 °C.<br />

(1650 to 1656 cm -1 at acidic pH) <strong>and</strong> a blue-shift <strong>of</strong> the amide<br />

II b<strong>and</strong> to 1542 cm -1 (1540 cm -1 for pH 2.5) is indicative <strong>of</strong> a<br />

certain increase <strong>of</strong> disordered structure (see Figure 6b). The<br />

appearance <strong>of</strong> a well-defined peak around 1625 cm -1 (1624<br />

cm -1 for pH 2.5) points to a structural transformation from an<br />

intramolecular hydrogen-bonded -sheet to an intermolecular<br />

hydrogen-bonded--sheet structure, 64 which is a structural<br />

characteristic <strong>of</strong> the amyloid fibrils. The spectrum also shows<br />

a high frequency component (∼1693 cm -1 ) that would suggest<br />

the presence <strong>of</strong> an antiparallel -sheet. 65 In addition, a small<br />

shoulder around 1534 cm -1 was also assigned to a -sheet. 66<br />

In the case <strong>of</strong> particulate gels formed under incubation at 65<br />

°C at pH 5.5, we could also observe a decrease <strong>and</strong> shift to<br />

1657 cm -1 <strong>of</strong> the amide I b<strong>and</strong> <strong>and</strong> an increase in the content<br />

<strong>of</strong> -sheet conformation. This is indicated by the enhancement<br />

<strong>of</strong> the intensity <strong>and</strong> further shift <strong>of</strong> the b<strong>and</strong> positioned around<br />

1626 cm -1 , in agreement with previous reports (see Figure<br />

6c). 61,67,68 In addition, the proportion <strong>of</strong> -sheet content is lower<br />

than that at pH 7.4 <strong>and</strong> 2.5. This corroborates that the<br />

aggregation near the protein isoelectric point takes place faster<br />

<strong>and</strong> nonspecifically which decreases the likehood <strong>of</strong> substantial<br />

structural rearrangements during the aggregation process. In<br />

contrast, amyloid fibrils <strong>and</strong> fibrillar gels resulted from partially<br />

highly charged unfolded states, 41,42 which involve long-range<br />

repulsion <strong>and</strong> slow aggregation occurring only when substantial<br />

structural reorganization allows the formation <strong>of</strong> a favorable<br />

structure, the cross- structure. In fact, when excess electrolyte<br />

is added, aggregation becomes faster <strong>and</strong> the proportion <strong>of</strong><br />

-sheet structure decreases due to an enhancement <strong>of</strong> the<br />

aggregation rates. 46<br />

Conclusions<br />

In this work, we have described the existence <strong>of</strong> suprafibrillar<br />

assemblies formed by the protein human serum<br />

181


12398 J. Phys. Chem. B, Vol. 113, No. 36, 2009 Juárez et al.<br />

albumin under different solution conditions: spherulites <strong>and</strong><br />

fibrillar gels. Under conditions where amyloid-like fibrils for<br />

this protein were previously detected, that is, upon incubation<br />

at 65 °C at both acidic <strong>and</strong> physiological pH in the presence<br />

<strong>of</strong> different amounts <strong>of</strong> added electrolyte, spherulite formation<br />

was detected. Within spherulites, fibrils display a radial<br />

arrangement around a disoganized protein core with sizes <strong>of</strong><br />

several micrometers, as revealed by POM <strong>and</strong> confocal<br />

microscopy. The spherulites are detected both in solution <strong>and</strong><br />

embedded in an isotropic matrix <strong>of</strong> a fibrillar gel. Upon<br />

extensive incubation under increasingly added electrolyte <strong>and</strong>/<br />

or protein concentration, gels are formed. Under conditions<br />

where the protein is electrically charged, fibrillar gels were<br />

observed. In contrast, at a pH close to the protein isolectric<br />

point, particulate gels were observed. Fibrillar gels are formed<br />

through intermolecular nonspecific association <strong>of</strong> amyloid<br />

fibrils at a pH far away from the isolectric point <strong>of</strong> the protein,<br />

as observed by TEM, AFM, <strong>and</strong> ATR-FTIR. For these types<br />

<strong>of</strong> gels, the protein molecules seem to display a “solid-like”<br />

behavior due to the existence <strong>of</strong> non-DLVO intermolecular<br />

repulsive forces, as observed by rheometry. As the solution<br />

ionic strength increases, a coarsening <strong>of</strong> this type <strong>of</strong> gels is<br />

observed as seen by AFM, with a decrease in the elastic<br />

response. In contrast, at pH 5.5, particulate gels are present<br />

as a consequence <strong>of</strong> a faster protein aggregation which does<br />

not allow the necessary structural reorganization to enable<br />

the formation <strong>of</strong> more ordered structures such as fibrils, as<br />

observed by their shorter gelation times <strong>and</strong> the frequencydependent<br />

behavior <strong>of</strong> the loss modulus. On the other h<strong>and</strong>,<br />

the formation <strong>of</strong> particulate gels with HSA also confirms the<br />

possibility that this type <strong>of</strong> structure may be a generic<br />

property <strong>of</strong> the polypeptide chains under suitable solution<br />

conditions.<br />

Acknowledgment. Authors thank Ministerio de Educación<br />

y Ciencia by project MAT-2007-61604 <strong>and</strong> Xunta de Galicia<br />

for financial support.<br />

Supporting Information Available: XRD fibril diffraction<br />

pattern <strong>and</strong> TEM pictures <strong>of</strong> mature fibrils. This material is<br />

available free <strong>of</strong> charge via the Internet at http://pubs.acs.org.<br />

References <strong>and</strong> Notes<br />

(1) Hady, J.; Selkoe, D. Science 2002, 297, 353–356.<br />

(2) Dobson, C. M. Nature 2003, 426, 884–890.<br />

(3) Stefani, M.; Dobson, C. M. J. Mol. Med. 2003, 81, 678–699.<br />

(4) Lansbury, P. T., Jr. Nat. Med. 2004, 10, 13709–13715.<br />

(5) Mattson, M. P. Nature 2004, 430, 631–639.<br />

(6) Soto, C.; Estrada, L.; Castilla, J. Trends Biochem. Sci. 2006, 31,<br />

150–155.<br />

(7) Khurana, R.; Ionescu-Zanetti, C.; Pope, M.; Li, J.; Nelson, M.;<br />

Ramírez- Alvarado, L.; Regan, L.; Fink, A. L.; Carter, S. A. Biophys. J.<br />

2003, 85, 1135–1144.<br />

(8) Chiti, F.; Stefani, M.; Taddei, N.; Ramponi, G.; Dobson, C. M.<br />

Nature 2003, 424, 805–808.<br />

(9) Williams, A. D.; Portelius, E.; Kheterpal, I.; Guo, J. T.; Cook, K. D.;<br />

Xu, Y.; Wetzel, R. J. Mol. Biol. 2004, 335, 833–842.<br />

(10) Hortscansky, P.; Christopeit, T.; Schroeckh, V.; Fändrich, M.<br />

Protein Sci. 2005, 14, 2915–2918.<br />

(11) Lashuel, H. A.; LaBrenz, S. R.; Woo, L.; Serpell, L. C.; Kelly,<br />

J. W. J. Am. Chem. Soc. 2000, 122, 5262–5277.<br />

(12) Uversky, V. N.; Segel, D. J.; Doniach, S.; Fink, A. L. Proc. Natl.<br />

Acad. Sci. U.S.A. 1998, 95, 5480–5483.<br />

(13) Gosal, W. S.; Clark, A. H.; Ross-Murphy, S. B. Biomacromolecules<br />

2004, 2, 2048–2419.<br />

(14) Sipe, J. D. Annu. ReV. Biochem. 1992, 61, 947–975.<br />

(15) Manuelidis, L.; Fritch, W.; Xi, Y.-G. Science 1997, 277, 94–98.<br />

(16) Krebs, M. R. H.; Bromley, E. H. C.; Rogers, S. S.; Donald, A. M.<br />

Biophys. J. 2005, 88, 2013–2021.<br />

(17) Ban, T.; Morigaki, K.; Yagi, H.; Kawasaki, T.; Kobayashi, A.;<br />

Yuba, S.; Naiki, H.; Goto, Y. J. Biol. Chem. 2006, 281, 33677–33683.<br />

(18) Heijna, M. C. R.; Telen, M. J.; van Enckevort, W. J. P.; Vlieg, E.<br />

J. Phys. Chem. B 2007, 111, 1567–1573.<br />

(19) Gosal, W. S.; Ross-Murphy, S. B. Curr. Opin. Colloid Interface<br />

Sci. 2000, 5, 188–194.<br />

(20) Bromley, E. H. C.; Krebs, M. R. H.; Donald, A. M. Faraday<br />

Discuss. 2005, 128, 13–27.<br />

(21) Bassett, D. C. J. Macromol. Sci. Phys. 2003, 42, 227–256.<br />

(22) Magill, J. H. J. Mater. Sci. 2001, 36, 3143–3164.<br />

(23) Murray, S. B.; Neville, A. C. Int. J. Biol. Macromol. 1997, 20,<br />

123–130.<br />

(24) Murray, S. B.; Neville, A. C. Int. J. Biol. Macromol. 1998, 22,<br />

137–144.<br />

(25) Rill, R. L. Proc. Natl. Acad. Sci. U.S.A. 1986, 83, 342–346.<br />

(26) Fezoui, Y.; Hartley, D. M.; Walsh, D. M.; Selkoe, D. J.; Osterhout,<br />

J. J.; Teplow, D. B. Nat. Struct. Biol. 2000, 7, 1095–1099.<br />

(27) Aggeli, A.; Bell, M.; Carric, L. M.; Fishwick, C. W. G.; Harding,<br />

R.; Mawer, P. J.; Radford, S. E.; Strong, A. E.; Boden, N. J. Am. Chem.<br />

Soc. 2003, 125, 9619–9628.<br />

(28) Hamodrakas, S. J.; Hoenger, A.; Iconomidou, V. A. J. Struct. Biol.<br />

2004, 145, 226–235.<br />

(29) Lockwood, N. A.; van Tankeren, R.; Mayo, K. H. Biomacromolecules<br />

2002, 3, 1225–1232.<br />

(30) Westlind-Danielsson, A.; Arneup, G. Biochemistry 2001, 40, 14736–<br />

14743.<br />

(31) Domike, K. R.; Donald, A. M. Biomacromolecules 2007, 8, 3930–<br />

3937.<br />

(32) Castelletto, V.; Hamley, I. W. Biomacromolecules 2007, 8, 77–<br />

83.<br />

(33) Ruth, L.; Eisenberg, D.; Neufeld, E. F. Acta Crystallogr., Sect. D<br />

2000, 56, 524–528.<br />

(34) Krebs, M. R. H.; MacPhee, C. E.; Miller, A. F.; Dunlop, I. E.;<br />

Dobson, C. M.; Donald, A. M. Proc. Natl. Acad. Sci. U.S.A. 2004, 101,<br />

14420–14424.<br />

(35) Jin, L.-W.; Claborn, K. A.; Kurimoto, M.; Geday, M. A.; Maezawa,<br />

I.; Sohraby, F.; Estrada, M.; Kaminsky, W.; Kahr, B. Proc. Natl. Acad.<br />

Sci. U.S.A. 2003, 100, 15294–15298.<br />

(36) Taniyama, H.; Kitamura, A.; Kagawa, Y.; Hirayama, K.; yoshino,<br />

T.; Kamiya, S. Vet. Pathol. 2000, 37, 104–107.<br />

(37) Acebo, E.; Mayorga, M.; Val-Bernal, J. F. Pathology 1999, 31,<br />

8–11.<br />

(38) Clark, A. H. G.; Kavanagh, G. M.; Ross-Murphy, S. B. Food<br />

Hydrocolloids 2001, 15, 383–400.<br />

(39) de la Fuente, M. A.; Singh, H.; Hemar, Y. Trends Food Sci. Technol.<br />

2002, 13, 262–274.<br />

(40) Zhang, S. Nat. Biotechnol. 2003, 21, 1171–1178.<br />

(41) Taboada, P.; Barbosa, S.; Castro, E.; Mosquera, V. J. Phys. Chem.<br />

B 2006, 110, 20733–20736.<br />

(42) Juárez, J.; Taboada, P.; Mosquera, V. Biophys. J. 2009, 96, 2353–<br />

2370.<br />

(43) Pace, C. N.; Vajdos, F.; Fee, L.; Grimsley, G.; Gray, T. Protein<br />

Sci. 1995, 4, 2411–2423.<br />

(44) Bolder, S. G.; Hendrickx, H.; Sagis, L. M. C.; van der Linden, E.<br />

J. Agric. Food Chem. 2006, 54, 4229–4234.<br />

(45) Juárez, J.; Taboada, P.; Mosquera, V. J. Phys. Chem. B 2009, 113,<br />

10521–10529.<br />

(46) Cerda-Costa, N.; Esteras-Chopo, A.; Aviles, F. X.; Serrano, L.;<br />

Villegas, V. J. Mol. Biol. 2007, 366, 1351–1363.<br />

(47) Plakoutsi, G.; Bemporad, F.; Calamai, M.; Taddei, N.; Dobson,<br />

C. M.; Chiti, F. J. Mol. Biol. 2005, 351, 910–922.<br />

(48) LeVine, H., III. Protein Sci. 1993, 2, 404–410.<br />

(49) Krebs, M. R. H.; Bromley, E. H. C.; Donald, A. M. J. Struct. Biol.<br />

2004, 149, 30–37.<br />

(50) Tobitani, A.; Ross-Murphy, S. B. Macromolecules 1997, 30, 4845–<br />

4854.<br />

(51) Ikeda, S.; Nishinari, K. Food Hydrocolloids 2001, 15, 401–406.<br />

(52) Surel, O.; Famelart, M. H. J. Dairy Res. 2003, 70, 253–256.<br />

(53) Saguer, F.; Fort, N.; Alvarez, P. A.; Sedman, J.; Ismail, A. A. Food<br />

Hydrocolloids 2008, 22, 459–467.<br />

(54) Zhao, J.; Majumdar, B.; Schulz, M. F.; Bates, F. S.; Almdal, K.;<br />

Mortensen, K.; Hadjuk, D. A.; Gruner, S. M. Macromolecules 1996, 29,<br />

1204–1215.<br />

(55) Radford, S. E.; Gosal, W. S.; Platt, G. W. Biochim. Biophys. Acta<br />

2005, 1753, 51–63.<br />

(56) Meehan, S.; Berry, Y.; Luisa, B.; Dobson, C. M.; Carver, J. A.;<br />

MacPhee, C. E. J. Biol. Chem. 2004, 279, 3413–3419.<br />

(57) Holm, N. K.; Jespersen, S. K.; Thomassen, L. V.; Wolff, T. Y.;<br />

Sehgal, P.; Thomsen, L. A.; Christiansen, G.; Andersen, C. B.; Knudsen,<br />

A. D.; Otzen, D. E. Biochim. Biophys. Acta 2007, 1774, 1128–1138.<br />

(58) Ikeda, S.; Morris, V. J. Biomacromolecules. 2002, 3, 382–389.<br />

(59) Calamai, M.; Canale, C.; Relini, A.; Stefani, M.; Chiti, F.; Dobson,<br />

C. M. J. Mol. Biol. 2005, 346, 603–616.<br />

182


Additional Supra-<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> HSA J. Phys. Chem. B, Vol. 113, No. 36, 2009 12399<br />

(60) Le Bon, C.; Nicolai, T.; Dur<strong>and</strong>, D. Macromolecules 1999, 32,<br />

6120–6127.<br />

(61) Krebs, M. R. H.; Devlin, G. L.; Donald, A. M. Biophys. J. 2007,<br />

92, 1336–1342.<br />

(62) Susi, H.; Byler, D. M. Biochem. Biophys. Res. Commun. 1983, 115,<br />

391–397.<br />

(63) Lin, S.-Y.; Wei, Y.-S.; Li, M.-J.; Wang, S.-L. Eur. J. Pharm.<br />

Biopharm. 2004, 57, 457–464.<br />

(64) Casal, H. L.; Kohler, U.; Mantsch, H. H. Biochim. Biophys. Acta<br />

1988, 957, 11–20.<br />

(65) Fabian, H.; Choo, L.-P; Szendrei, G. I.; Jackson, M.; Halliday,<br />

W. C.; Otvos, L., Jr.; Mantsch, H. H. Appl. Spectrosc. 1993, 47, 1513–<br />

1518.<br />

(66) Lin, S.-Y.; Wei, Y.-S.; Li, M.-J.; Wang, S.-L. Spectrochim. Acta,<br />

Part A 2004, 60, 3107–3111.<br />

(67) Ikeda, S.; Li-Chan, E. C. Y. Food Hydrocolloids 2004, 18, 489–<br />

498.<br />

(68) Ikeda, S.; Nishinari, K. Biopolymers 2001, 59, 87–102.<br />

JP904167E<br />

183


5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

45<br />

S<strong>of</strong>t Matter<br />

Cite this: DOI: 10.1039/c0xx00000x<br />

www.rsc.org/xxxxxx<br />

Dynamic Article Links ►<br />

PAPER<br />

Hydration effects on the fibrillation process <strong>of</strong> a globular protein: The<br />

case <strong>of</strong> human serum albumin.<br />

Josué Juárez, Manuel Alatorre-Meda, Adriana Cambón, Antonio Topete, Silvia Barbosa, Pablo<br />

Taboada*, <strong>and</strong> Víctor Mosquera.<br />

Received (in XXX, XXX) Xth XXXXXXXXX 200X, Accepted Xth XXXXXXXXX 200X<br />

DOI: 10.1039/b000000x<br />

In this work, we have studied the fibrillation process <strong>of</strong> human serum albumin (HSA) under different<br />

solution conditions. In particular, aggregation kinetics, fibril morphology, <strong>and</strong> composition structural<br />

changes were investigated at varying experimental conditions such as pH (2.0 <strong>and</strong> 7.4), temperature (at 25<br />

<strong>and</strong> 65 ºC), <strong>and</strong> solvent polarity (ethanol/water mixtures, 10 – 90% v/v). The characterization was carried<br />

out by means <strong>of</strong> static <strong>and</strong> dynamic light scattering (SLS <strong>and</strong> DLS), ThT fluorescence, circular dichroism<br />

(CD) <strong>and</strong> Fourier Transform Infrared spectroscopies (FT-IR), <strong>and</strong> transmission electron microscopy<br />

(TEM). The aggregation process <strong>and</strong> the α-helix to β-sheet transitions were found to be favored by<br />

temperature <strong>and</strong> physiological pH. Also, pH was observed to influence both the fibrillation pathway <strong>and</strong><br />

aggregation kinetics, changing from a classical fibrillation process with a lag phase under acidic<br />

conditions to a downhill polymerization process at physiological pH in the presence <strong>of</strong> the alcohol.<br />

Regarding protein structural composition, at room temperature <strong>and</strong> physiological pH ethanol was found to<br />

promote an α-helix to β-sheet conformational transition at intermediate alcohol concentrations, whereas at<br />

low <strong>and</strong> high ethanol contents α-helix prevailed as the predominant structure. Under acidic conditions,<br />

ethanol promotes an important fibrillation at high cosolvent concentrations due to screening <strong>of</strong> electric<br />

charges <strong>and</strong> a decrease in solvent polarity. On the other h<strong>and</strong>, important differences in the morphology <strong>of</strong><br />

the resulting fibrils <strong>and</strong> aggregates are observed depending on the solution conditions. In particular, the<br />

formation <strong>of</strong> classical amyloid-like fibrils at physiological pH <strong>and</strong> high temperature are observed, in<br />

contrast to the usual curly morphology displayed by HSA fibrils under most <strong>of</strong> solution conditions.<br />

Although high temperature <strong>and</strong> pH are the main parameters influencing the protein structure<br />

destabilization <strong>and</strong> subsequent aggregation upon incubation, ethanol helps to regulate the hydrogen<br />

bonding, the attractive hydrophobic interactions, <strong>and</strong> the protein accessible surface area, thus, modifying<br />

packing constraints <strong>and</strong> the resulting aggregate morphologies.<br />

Introduction<br />

From a general perspective, amyloid fibril formation present<br />

pros <strong>and</strong> cons regarding different approaches such as human<br />

health or technological development. On one h<strong>and</strong>, amyloid<br />

fibrils are recognized as high-performance protein nanomaterials<br />

with a formidable rigidity;<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 1<br />

1-2 on the other, amyloidosis, the<br />

clinical condition in which amyloid fibrils form from innocuous<br />

soluble proteins, has been found to be involved in diverse<br />

pathological responses as observed in various neurodegenerative<br />

disorders including Alzheimer’s, Parkinson’s, <strong>and</strong> prion diseases<br />

amongst others. 3-5 applications in the area <strong>of</strong> nanobiotechnology.<br />

Concerning protein folding <strong>and</strong> its repercussion on amyloid<br />

fibril formation, it has been demonstrated that fibrillation may<br />

proceed from different protein conformations including<br />

completely/partially unfolded<br />

The elucidation <strong>of</strong> the underlying molecular<br />

mechanisms for fibril formation would result not only in the<br />

development <strong>of</strong> strategies for the treatment <strong>of</strong> amyloidosis-related<br />

disorders but also in the design <strong>of</strong> new biomaterials for future<br />

Grupo de Física de Coloides y Polímeros, Departamento de Física de la<br />

Materia Condensada, Facultad de Física, Universidad de Santiago de<br />

Compostela, E-15782. Santiago de Compostela. Spain.<br />

pablo.taboada@usc.es<br />

6-11 or fully folded states, 12,13<br />

55<br />

although the essential nature <strong>of</strong> the resulting amyloid fibrils can<br />

be largely independent <strong>of</strong> the conformational properties <strong>of</strong> the<br />

soluble precursors. 6<br />

In living organisms, the native environment <strong>of</strong> proteins is a<br />

60 complex composition <strong>of</strong> water, cosolvents <strong>and</strong> cosolutes which<br />

affect the stability <strong>of</strong> the native protein fold. 14 In this way,<br />

different experimental approaches using the addition <strong>of</strong><br />

cosolvents <strong>and</strong> cosolutes to mimic various cellular environments<br />

in vitro have provided further knowledge on protein folding <strong>and</strong><br />

aggregation/fibrillation mechanisms. 15-20 65<br />

Thereby, cosolventmediated<br />

systems have revealed that differences in solvent<br />

185<br />

50


properties induce solvent-adapted structural “responses” <strong>of</strong> the<br />

aggregating protein, which leads to a structural diversity <strong>of</strong> the<br />

resulting aggregates/fibers. 21-23 In particular, alcohols are<br />

included among the most commonly studied cosolvents. Several<br />

studies have shown that the effects <strong>of</strong> alcohol on very different<br />

systems <strong>and</strong> processes such as the thermal denaturation <strong>of</strong> nucleic<br />

acids <strong>and</strong> proteins, protein folding, micellization <strong>of</strong> surfactant<br />

molecules, or the solubility <strong>of</strong> nonelectrolytes are strikingly<br />

similar. 24-26 Also, water/alcohol mixtures at moderately low pH<br />

values have been suggested as model systems for studying the<br />

joint action <strong>of</strong> the local decrease in both pH <strong>and</strong> dielectric<br />

constant on the protein structure near the membrane surface; 27-32<br />

similarly, new ways <strong>of</strong> protein administration <strong>and</strong> delivery based<br />

on inhalation systems with non-aqueous solvents as suspension<br />

media also reinforce the necessity <strong>of</strong> clarifying how the presence<br />

<strong>of</strong> the solvent affects the protein conformation <strong>and</strong> biological<br />

24, 25<br />

activity.<br />

From a thermodynamic st<strong>and</strong>point, the protein accessible<br />

surface area (ASA), the solvent exposure <strong>of</strong> hydrophobic<br />

residues, <strong>and</strong> a solvophobic backbone 33 can be taken as barriers<br />

hampering protein-solvent interactions. 34 Therefore, the<br />

hierarchical assembly <strong>of</strong> amyloids can perfectly be understood as<br />

an alternative to the native packing conformational struggle <strong>of</strong> a<br />

polypeptide chain, reducing ASA <strong>and</strong> saturating hydrogen<br />

bonding, while a simultaneous decrease in the configurational<br />

entropy <strong>of</strong> the protein is <strong>of</strong>fset by gains in solvent entropy. The<br />

role <strong>of</strong> hydrational forces in protein aggreation in vivo, while still<br />

poorly understood, may be instrumental in elucidating the<br />

molecular basis <strong>of</strong> amyloidosis <strong>and</strong>, as such, attracts considerable<br />

interest. 35 In this way, the propensity for amyloid formation in<br />

mixed solvents <strong>of</strong> different proteins as, for example, human<br />

serum albumin (HSA), hen egg white lysozyme, bovine βlactoglobulin,<br />

histone H2A, insulin, <strong>and</strong> bovine trypsinogen<br />

amongst many others, has been characterized. 36-39 Amongst them,<br />

HSA has been proposed as a good model for protein aggregation<br />

studies 40-42 as a consequence <strong>of</strong> both its physiological<br />

implications as a carrier protein <strong>and</strong> blood pressure regulator,<br />

together with its propensity to easily aggregate in vitro. Previous<br />

reports have addressed the HSA fibrillation kinetics, fibrillation<br />

pathway, <strong>and</strong> intermediate <strong>and</strong> final protein aggregated structures<br />

under varying pH <strong>and</strong> ionic strength solution conditions. 40-42<br />

However, a detailed characterization <strong>of</strong> the aggregation/fibril<br />

process <strong>and</strong> the structure <strong>of</strong> the resulting aggregates in mixed<br />

solvent solutions under varying external conditions is still lacked.<br />

In the present work, we present a detailed study on the<br />

aggregation/fibrillation pathway <strong>of</strong> the protein human serum<br />

albumin (HSA) in water/ethanol mixed solvent solutions at<br />

varying pH <strong>and</strong> temperature conditions. This enables us to make<br />

a clear comparison between the fibrillation pathway <strong>and</strong><br />

intermediate/final protein aggregates in the mixed solvent with<br />

those obtained in pure aqueous solution in order to define the role<br />

<strong>of</strong> hydrational forces on the protein fibrillation mechanism. In<br />

fact, changes on the fibrillation mechanism from a downhill to a<br />

nucleated-growth polymerization mechanism at acidic pH <strong>and</strong><br />

high temperature solution conditions are confirmed. Overall, as a<br />

result <strong>of</strong> differences in packing depending on the solution<br />

condition, different structures for the resulting fibrils are<br />

observed, from classically thin <strong>and</strong> very long straight to shorter<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

45<br />

50<br />

55<br />

<strong>and</strong> more curly fibers. In this regard, it comes as no surprise that<br />

factors favoring or disfavoring exposure <strong>of</strong> solvophobic groups<br />

must affect thermodynamic preferences for particular<br />

conformations.<br />

2 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society <strong>of</strong> Chemistry [year]<br />

186<br />

60<br />

65<br />

Materials <strong>and</strong> Methods<br />

Materials.<br />

Human serum albumin (70024-90-7), <strong>and</strong> Thi<strong>of</strong>lavin T (ThT)<br />

were obtained from Sigma Chemical Co. ThT was used as<br />

70 received. All other chemicals were <strong>of</strong> the highest purity available.<br />

75<br />

80<br />

85<br />

90<br />

95<br />

100<br />

105<br />

110<br />

115<br />

Preparation <strong>of</strong> HSA solutions.<br />

HSA was purified by liquid chromatography using a<br />

Superdex 75 column equilibrated with 0.01 M phosphate buffer<br />

before use. To prepare the protein solutions, increasing volumes<br />

<strong>of</strong> ethanol were added to 1.0 mL <strong>of</strong> HSA stock solution either at<br />

pH 7.4 (sodium monophosphate-sodium diphosphate buffer) or<br />

pH 2.0 (glycine + HCl buffer), to get the desired ethanol<br />

concentration in the mixed solvent. The final protein<br />

concentration in all cases was 10 mg/mL with ionic strength 10<br />

mM. pH was kept constant by adding HCl or NaOH when<br />

needed. The protein concentration was determined<br />

spectrophotometrically, using a molar absorption coefficient <strong>of</strong><br />

35219 M -1 cm -1 at 280 nm. 43 Aliquots were added through an<br />

electronic disperser unit Dosimat Metrohm 765. Experiments<br />

were carried out using double distilled, deionized <strong>and</strong> degassed<br />

water. Before incubation, solutions were filtered through a 0.2<br />

µm filter into sterile test tubes. Samples were incubated at a<br />

specified temperature in a refluxed reactor for 15-20 days as<br />

required.<br />

Light Scattering.<br />

DLS <strong>and</strong> SLS intensities were measured at 25 °C by means <strong>of</strong><br />

an ALV-5000F (ALV-GmbH) instrument working with a<br />

vertically polarized incident light <strong>of</strong> wavelength λ = 488 nm<br />

supplied by a CW diode-pumped Nd; YAG solid-state laser<br />

supplied by Coherent, Inc. <strong>and</strong> operated at 400 mW. The intensity<br />

scale was calibrated against scattering from toluene.<br />

Measurements were carried out at a scattering angle θ = 90° to<br />

the incident beam. Solutions were equilibrated for 30 min before<br />

measurements to reach thermal stabilization. Experiment duration<br />

was in the range <strong>of</strong> 3-5 min, <strong>and</strong> each experiment was repeated<br />

two or more times. The correlation functions from DLS were<br />

analyzed using the CONTIN method to obtain intensity<br />

distributions <strong>of</strong> decay rates (Γ). 44 The decay rate distribution<br />

functions gave distributions <strong>of</strong> apparent diffusion coefficient<br />

(D app = Γ /q 2 ,where the scattering vector q = (4πns/λ)sin(θ /2),<br />

<strong>and</strong> ns is the refractive index <strong>of</strong> solvent) <strong>and</strong> integrating over the<br />

intensity distribution gave the intensity-weighted average <strong>of</strong> D . app<br />

Values <strong>of</strong> the apparent hydrodynamic radius (r , radius <strong>of</strong><br />

h,app<br />

hydrodynamically equivalent hard sphere corresponding to D ) app<br />

were calculated from the Stokes-Einstein equation:


5<br />

10<br />

15<br />

20<br />

25<br />

r h,app = kT/(6πηD app ) (1)<br />

where k is the Boltzmann constant <strong>and</strong> η is the viscosity <strong>of</strong> water<br />

at temperature T.<br />

Thi<strong>of</strong>lavin T spectroscopy.<br />

Protein <strong>and</strong> ThT were dissolved in buffer at a final proteindye<br />

molar ratio <strong>of</strong> 50:1. Samples were continuously stirred only<br />

during data acquisition intervals in order to avoid ThT deposition<br />

onto protein aggregates/fibrils. Fluorescence was measured in a<br />

Cary Eclipse fluorescence spectrophotometer equipped with a<br />

temperature controller <strong>and</strong> a multi-cell sample holder (Varian<br />

Instruments Inc.). Excitation <strong>and</strong> emission were at wavelengths <strong>of</strong><br />

450 <strong>and</strong> 482 nm, respectively. All intensities were backgroundcorrected<br />

for the ThT-fluorescence in the respective solvent in the<br />

absence <strong>and</strong> presence <strong>of</strong> monomeric HSA at pH 7.4 <strong>and</strong> 2.0.<br />

Circular Dichroism (CD).<br />

Far-UV circular dichroism (CD) spectra were obtained using<br />

a JASCO-715 automatic recording spectropolarimeter with a<br />

JASCO PTC-343 Peltier-type thermostated cell holder. Quartz<br />

cuvettes with 0.2 cm pathlength were used. The protein<br />

concentration was reduced to 0.1 mg/mL in order to avoid light<br />

scattering effects as much as possible due to protein aggregation.<br />

CD spectra were obtained from aliquots withdrawn from the<br />

aggregation mixtures at the indicated conditions after incubation,<br />

<strong>and</strong> recorded at wavelengths between 195 <strong>and</strong> 300 nm at 25 ºC.<br />

The mean residue ellipticity θ (deg cm 2 dmol -1 30<br />

) was calculated<br />

35<br />

40<br />

45<br />

50<br />

55<br />

according to: θ=(θObs/10)/(MRM/lc), where θobs is the<br />

observed ellipticity (in deg), MRM is the mean residue molecular<br />

mass (in g mol -1 ), l is the optical path-length (in cm), <strong>and</strong> c is the<br />

protein concentration (in g ml -1 ). The secondary structure <strong>of</strong> the<br />

protein was calculated from far-UV CD spectra by running the<br />

SELCON3, CONTIN, <strong>and</strong> DSST programs within the<br />

DICHROWEB server. 45,46 Final results were assumed when data<br />

generated from all programs showed convergence.<br />

Transmission electron microscopy (TEM).<br />

HSA solutions were applied to carbon-coated copper grids,<br />

blotted, washed, negatively stained with 2% (w/v) <strong>of</strong><br />

phosphotungstic acid, air dried, <strong>and</strong> then examined with a Phillips<br />

CM-12 transmission electron microscope operating at an<br />

accelerating voltage <strong>of</strong> 120 kV. Samples were diluted between<br />

20-200-fold prior deposition onto the grids.<br />

Results <strong>and</strong> Discussion<br />

To induce HSA aggregation <strong>and</strong> subsequent amyloid<br />

formation we use 10 mg/mL HSA solutions at either pH 7.4 or<br />

pH 2.0 <strong>and</strong> ionic strength <strong>of</strong> 10 mM. At pH 7.4 the protein is in<br />

its native state whereas at pH 2.0 it is in its E-exp<strong>and</strong>ed one,<br />

which is characterized by a separation <strong>of</strong> domain III <strong>and</strong> a loss <strong>of</strong><br />

the intradomain helices <strong>of</strong> domain I. 47 Monomer sizes <strong>and</strong><br />

structural compositions were similar to those previously<br />

reported. 40 Samples were prepared by adding specified volumes<br />

<strong>of</strong> protein stock solutions <strong>and</strong> different amounts <strong>of</strong> ethanol, which<br />

has been shown to destabilize both the secondary <strong>and</strong> tertiary<br />

structures <strong>of</strong> HSA; 36 subsequently, samples were incubated at<br />

room or elevated temperature (65 ºC, above the melting<br />

temperature <strong>of</strong> HSA) 48,49 to induce further protein structural<br />

changes <strong>and</strong> aggregation.<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 3<br />

187<br />

60<br />

65<br />

70<br />

75<br />

80<br />

85<br />

90<br />

95<br />

100<br />

105<br />

110<br />

Aggregate formation.<br />

The formation <strong>of</strong> HSA aggregates <strong>and</strong> their aggregation<br />

kinetics were firstly analyzed by SLS. Figure 1 shows the time<br />

evolution <strong>of</strong> the scattered light intensity <strong>of</strong> HSA samples at<br />

varying pH, temperature <strong>and</strong> alcohol content conditions upon<br />

incubation. Different curve pr<strong>of</strong>iles are obtained depending on<br />

incubation conditions.<br />

At physiological pH (Figure 1a-b), the plots reflect a quick<br />

increase in the scattered intensity at short incubation times, which<br />

denotes the fast formation <strong>of</strong> protein aggregates. SLS plot pr<strong>of</strong>iles<br />

suggest a mechanism <strong>of</strong> continuous protein aggregation (in<br />

particular, a continuous fibrillation process as later confirmed by<br />

ThT, CD <strong>and</strong> TEM data) without a nucleation step, as denoted the<br />

absence <strong>of</strong> a lag phase in the early periods <strong>of</strong> incubation (see inset<br />

in Figure 1a). In addition, the absence <strong>of</strong> a faster aggregation<br />

process when protein seeds (i.e. protein aggregates) are added to<br />

protein solutions followed by subsequent incubation, <strong>and</strong> the<br />

decrease in the aggregation rate if the starting protein<br />

concentration is lowered (see Figure S1 in ESI) further indicate<br />

that under the present conditions HSA aggregation occurs by<br />

means <strong>of</strong> a classical coagulation or downhill polymerization<br />

mechanism, i.e. each protein monomer association is bimolecular,<br />

effectively irreversible <strong>and</strong> thermodynamically favorable, so that<br />

there is no energy barrier to impede aggregate growth. 50 A similar<br />

behavior has been previously obtained in the absence <strong>of</strong><br />

alcohol. 40-42,50,51 ThT fluorescence data confirmed similar trends<br />

as those observed by SLS (see Figure S2). High values <strong>of</strong> ThT<br />

fluorescence point to an important formation/gain <strong>of</strong> β-sheet<br />

structure along the protein aggregation process, which is a first<br />

evidence <strong>of</strong> the possible formation <strong>of</strong> HSA fibrils during the<br />

aggregation process, 37 as further confirmed below by CD <strong>and</strong><br />

TEM techniques.<br />

Also, under physiological conditions we observe that as the<br />

ethanol concentration increases up to 60% (v/v) the formation <strong>of</strong><br />

more numerous <strong>and</strong>/or bulkier protein aggregates takes place as<br />

derived from both the increments in scattered intensities –<br />

proportional to protein concentration <strong>and</strong> molecular weight<br />

(Figure 1a,b), <strong>and</strong> the shifts <strong>of</strong> the aggregate size distributions to<br />

larger values (see Figure 2a-b below). The clustering <strong>of</strong> alcohol<br />

molecules <strong>and</strong> the enhancement <strong>of</strong> alcohol-water interactions at<br />

intermediate ethanol concentrations in the mixed solvent should<br />

modify the extent <strong>of</strong> protein hydration <strong>and</strong> molecular packing 53 as<br />

confirmed by CD <strong>and</strong> FT-IR below, favoring an enhanced protein<br />

aggregation. At larger alcohol concentrations (> 60% (v/v)),<br />

protein solutions became progressively turbid as a consequence<br />

<strong>of</strong> the formation <strong>of</strong> even larger aggregates, 52 which gradually<br />

precipitated (see below for further explanation).


5<br />

10<br />

15<br />

20<br />

25<br />

S<strong>of</strong>t Matter<br />

Cite this: DOI: 10.1039/c0xx00000x<br />

www.rsc.org/xxxxxx<br />

Intensity (a.u.)<br />

Intensity (a.u.)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

280<br />

240<br />

200<br />

160<br />

120<br />

80<br />

40<br />

0<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000<br />

Time (min)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Intensity (a.u.)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 500 1000 1500 2000 2500 3000<br />

Time (min)<br />

0 2000 4000 6000<br />

Time (min)<br />

8000 10000 12000<br />

Dynamic Article Links ►<br />

PAPER<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 4<br />

188<br />

a)<br />

c)<br />

Intensity (a.u)<br />

Intensity (a.u.)<br />

1200<br />

1100<br />

1000<br />

900<br />

800<br />

100<br />

50<br />

0<br />

24<br />

20<br />

16<br />

12<br />

8<br />

4<br />

Intensity (a.u.)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0 5 10 15 20 25 30<br />

Time (min)<br />

0 1000 2000 3000 4000 5000 6000 7000 8000 9000<br />

Time (min)<br />

0<br />

0 5000 10000 15000 20000 25000<br />

Time (min)<br />

Figure 1: Time evolution <strong>of</strong> the scattered intensity <strong>of</strong> HSA solutions at a) pH 7.4 <strong>and</strong> 65 ºC, b) pH 7.4 <strong>and</strong> 25 ºC, c) pH 2.0 <strong>and</strong> 65 ºC,<br />

<strong>and</strong> d) pH 2.0 <strong>and</strong> 25 ºC under different concentrations <strong>of</strong> ethanol in the mixed solvent. In a), b) <strong>and</strong> d) ethanol concentrations are ()<br />

20, () 40, () 60, <strong>and</strong> () 80% (v/v). The enhanced scattered dispersion above 1000 a.u. at high alcohol concentrations in the mixed<br />

solvent is a consequence <strong>of</strong> the formation <strong>of</strong> big aggregates. In c) ethanol concentrations in the mixed solvent are () 20, () 40, ()<br />

50, () 60, <strong>and</strong> () 80% (v/v). The inset in Figure 1b illustrates, as an example, the absence <strong>of</strong> a lag phase in the incubation <strong>of</strong> HSA at<br />

25 ºC in the presence <strong>of</strong> 20% (v/v). The inset in Figure 1c illustrates the existence <strong>of</strong> the first step in the aggregation kinetics.<br />

For samples incubated under acidic conditions (Figure 1c-d),<br />

the scattered intensity plot pr<strong>of</strong>iles depend on both alcohol<br />

content <strong>and</strong> temperature. HSA samples incubated at 25 ºC in the<br />

whole range <strong>of</strong> mixed solvent compositions <strong>and</strong> those incubated<br />

at 65 ºC at ethanol concentrations below 50% (v/v) gave rise to<br />

SLS plots with classical sigmoidal pr<strong>of</strong>ile, i.e., there exists an<br />

initial lag phase in which the scattered intensity remains almost<br />

constant followed by a subsequent exponential phase which ends<br />

up in a final equilibrium region. These three stages correspond to<br />

the nucleation, elongation, <strong>and</strong> equilibrium phases. 54,55 The<br />

presence <strong>of</strong> a lag phase at acidic pH is related to the different<br />

initial structure <strong>of</strong> protein molecules in acidic solution, <strong>and</strong> to<br />

changes in the nature <strong>and</strong> strength <strong>of</strong> intermolecular interactions<br />

upon incubation, as previously stated in previous works. 41,42 This<br />

finding is additionally confirmed when a seeding growth process<br />

is performed, in which the lag phase is abolished (see Figure<br />

S3a).<br />

On the other h<strong>and</strong>, for HSA samples at 65 ºC in the presence<br />

<strong>of</strong> alcohol concentrations between 50 <strong>and</strong> 90% (v/v), the<br />

formation <strong>of</strong> protein aggregates occurs in two well-differentiated<br />

30<br />

35<br />

40<br />

45<br />

steps: there exists a first short increase in the scattered intensity at<br />

very low incubation times followed by a quasi-plateau region;<br />

then, a second much stepper rise in a narrow time interval takes<br />

place until the final equilibrium region is reached. It has been<br />

previously demonstrated that this plot pr<strong>of</strong>ile emerges when small<br />

oligomeric structures rapidly form, 40 which need more time to<br />

develop <strong>and</strong>/or persist for longer times due to their enhanced<br />

solubility under acidic conditions. 40-41 These species are watersoluble<br />

<strong>and</strong>, hence, only after achieving a critical<br />

concentration/size the energy l<strong>and</strong>scape <strong>of</strong> the solution changes,<br />

i.e., the aggregation process becomes thermodynamically<br />

favorable enabling the evolvement <strong>of</strong> small oligomers to larger<br />

aggregates 56 as confirmed below by DLS data. Hence, the quasiplateau<br />

region has been identified as a lag phase, as confirmed its<br />

suppression when a seeding fibril growth process is performed<br />

under the present solution conditions (see Figure S3b). It is also<br />

worth mentioning that the scattered intensities at equilibrium<br />

increases at ethanol concentrations between 0 <strong>and</strong> 50% (v/v),<br />

followed by a decrease between 50-70% (v/v), <strong>and</strong> by a new final<br />

increment (see Figure 1c). As confirmed by TEM <strong>and</strong> CD data<br />

b)<br />

d)


5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

45<br />

50<br />

below, the first increase is related to the presence <strong>of</strong> progressively<br />

interconnected aggregates from the reduction in solvent polarity,<br />

which originates subtle changes in the protein secondary structure<br />

composition favoring the entanglement <strong>of</strong> protein aggregates. The<br />

subsequent intensity decrease at larger alcohol concentrations (><br />

50% v/v) originates from the formation <strong>of</strong> well-dispersed<br />

aggregates (short fibrils, see TEM pictures below) as a<br />

consequence <strong>of</strong> additional changes in protein secondary structure<br />

(as confirmed CD data below). The final scattered intensity<br />

increase at the largest ethanol contents arises from the presence <strong>of</strong><br />

more numerous <strong>and</strong> longer aggregated structures in solution.<br />

These observations are in contrast with the behavior observed<br />

at 25 ºC (see Figure 1d). For samples containing low ethanol<br />

contents, their low scattering intensity suggests that almost no<br />

aggregation occurs, whereas at higher ethanol contents (>40%<br />

v/v) the rise in the scattered intensity pr<strong>of</strong>ile corroborates the<br />

presence <strong>of</strong> aggregates as a consequence <strong>of</strong> the screening <strong>of</strong><br />

electrostatic repulsions between protein molecules as the solvent<br />

permittivity decreases. Additional confirmation is given by CD<br />

<strong>and</strong> ThT fluorescence data below.<br />

On the other h<strong>and</strong>, comparison <strong>of</strong> scattered intensities at<br />

acidic <strong>and</strong> physiological pH shows the scattered intensity at the<br />

former pH is, in general, lower than that at the latter at the same<br />

mixed solvent conditions (see Figure 1). This is a consequence <strong>of</strong><br />

the lowering <strong>of</strong> intermolecular interactions due to protonation <strong>of</strong><br />

positively charged residues. 57 Also, when comparing the effect <strong>of</strong><br />

temperature on the scattered intensity pr<strong>of</strong>iles, we can observe<br />

that incubation at 25 ºC involves a lower scattered intensity at<br />

both pH in the whole range <strong>of</strong> solvent compositions. It is known<br />

that hydrogen bonding is weakened as temperature rises, but<br />

hydrophobic interactions become strengthened. 58 As the heating<br />

process proceeds, destabilization <strong>of</strong> the protein helical structure<br />

<strong>of</strong> HSA occurs though hydrogen bonding weakening, as denoted<br />

by the loss <strong>of</strong> the minima at 208 <strong>and</strong> 222 nm typically assigned to<br />

α-helices in the HSA native state in the CD pr<strong>of</strong>iles <strong>of</strong> HSA<br />

samples at high temperature (see Figure 3); then, previously<br />

buried aminoacid residues can be now exposed to solvent as<br />

follows from changes in protein conformation from CD data (see<br />

below), which results more prone to aggregation as observed<br />

from the SLS data (<strong>and</strong> further confirmed by TEM pictures<br />

below).<br />

Aggregation kinetics.<br />

Data analysis <strong>of</strong> scattered intensity plots with a sigmoidal-like<br />

pr<strong>of</strong>ile is consistent (also according to spectroscopy <strong>and</strong> TEM<br />

data), at a first approximation, with the following kinetic scheme:<br />

M ↔ I ↔ nucleus → fibrils<br />

where M is the monomer <strong>and</strong> I is the intermediate. Thus, the<br />

scattered intensity as a function <strong>of</strong> time was fitted to the<br />

following equation:<br />

(2)<br />

where Y is the scattered intensity, x the time, <strong>and</strong> xo is the time to<br />

reach 50% <strong>of</strong> maximal scattered intensity. Thus, the apparent rate<br />

constant, kapp, <strong>of</strong> fibril growth is given by 1/τ, <strong>and</strong> the lag time by<br />

xo - 2τ. 59 Lag times <strong>and</strong> apparent first-order rate constants were<br />

determined from curve fits <strong>and</strong> shown in Table S1 in ESI. In<br />

general, an increase in the alcohol concentration resulted in<br />

longer lag times <strong>and</strong> larger apparent rate constants. Also, the<br />

aggregation kinetics at physiological pH appears to be faster than<br />

under acidic conditions. This behavior is probably related to the<br />

existence <strong>of</strong> electrostatic repulsions between protein molecules<br />

<strong>and</strong> to an enhanced solubility <strong>of</strong> the initial protein clusters under<br />

acidic conditions. Finally, it is necessary to mention that two<br />

growth rates were obtained at high ethanol concentrations upon<br />

incubation at acidic conditions <strong>and</strong> 25 ºC. This fact is originated<br />

by the presence <strong>of</strong> intermediate oligomeric structures along the<br />

HSA aggregation pathway under these solution conditions, as<br />

discussed elsewhere. 40<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 5<br />

55<br />

60<br />

65<br />

70<br />

75<br />

80<br />

85<br />

90<br />

95<br />

100<br />

105<br />

Aggregate size distribution.<br />

Figure 2 shows some examples <strong>of</strong> population size<br />

distributions for samples incubated under the different solution<br />

conditions. At 65 ºC <strong>and</strong> physiological pH, two different<br />

populations are well differentiated <strong>and</strong> present under all mixed<br />

solvent compositions (Figure 2a): A peak at large sizes (ca. 400-<br />

1000 nm) assigned to protein aggregates, <strong>and</strong> other at smaller<br />

sizes (at ca. 20 to 40 nm) probably representing protein<br />

clusters/oligomers. Also, it is observed that a third population <strong>of</strong><br />

aggregates with very large sizes (∼ 20 µm) appears at an ethanol<br />

concentration <strong>of</strong> 80% (v/v). This feature confirms that supraaggregation<br />

<strong>of</strong> protein aggregates is favored at high alcohol<br />

contents under the present conditions, in agreement with SLS<br />

data, <strong>and</strong> as further confirmed below by TEM pictures. For<br />

samples incubated at 25 ºC <strong>and</strong> physiological pH <strong>and</strong> at 65 ºC <strong>and</strong><br />

pH 2.0 (Figure 2b-c), a progressive increase in the size <strong>of</strong> protein<br />

aggregates takes place as the ethanol concentration increases, as<br />

denoted by the existence <strong>of</strong> bimodal distributions. Only at<br />

intermediate ethanol concentrations the referred size increase<br />

involves the superimposition <strong>of</strong> population sizes corresponding to<br />

large protein aggregates <strong>and</strong> oligomeric structures, creating a<br />

wide population distribution (but still bimodal, as observed after<br />

deconvolution). In contrast, incubation at pH 2.0 <strong>and</strong> 25 ºC<br />

results in the simultaneous decrease <strong>of</strong> the peak height<br />

corresponding to the smallest sizes <strong>and</strong> the increase <strong>of</strong> that<br />

corresponding to the biggest ones. This fact indicates a<br />

progressive evolvement <strong>of</strong> protein monomers <strong>and</strong> small<br />

oligomeric structures to aggregates <strong>of</strong> larger sizes as the ethanol<br />

concentration in the mixed solvent increases (Figure 2d),<br />

corroborating SLS data.<br />

Regarding the hydrodynamic radii derived from the size<br />

distributions depicted in Figure 2a-d, they can be grouped into<br />

three average values corresponding to rh,app <strong>of</strong> ca. 400-1000 nm,<br />

20-50 nm, <strong>and</strong> 2-5 nm. At this point, it is necessary to remind that<br />

DLS sizes are obtained assuming a spherical shape, so these<br />

189


5<br />

10<br />

15<br />

S<strong>of</strong>t Matter<br />

Cite this: DOI: 10.1039/c0xx00000x<br />

www.rsc.org/xxxxxx<br />

Intensity (a.u.)<br />

Intensity (a.u.)<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1 10 100 1000 10000 100000<br />

r (nm)<br />

h,app<br />

1 10 100 1000 10000 100000<br />

r (nm)<br />

h,app<br />

r h,app (nm)<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

e)<br />

a)<br />

c)<br />

Dynamic Article Links ►<br />

0.1 1 10 100<br />

r (nm)<br />

h,app<br />

1000 10000 100000<br />

PAPER<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 6<br />

190<br />

Intensity (a.u.)<br />

Intensity (a.u.)<br />

0<br />

10 20 30 40 50 60 70 80 90 100<br />

% (v/v) ethanol<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.1 1 10 100<br />

r (nm)<br />

h,app<br />

1000 10000 100000<br />

Figure 2: Selected size distributions derived from DLS data for HSA solutions incubated at pH 7.4 <strong>and</strong> a) 65ºC, <strong>and</strong> b) 25 ºC at ethanol<br />

concentrations <strong>of</strong> (⎯) 20, (⎯) 60, <strong>and</strong> (⎯) 80% (v/v), respectively; <strong>and</strong> at pH 2.0 <strong>and</strong> c) 65 ºC <strong>and</strong> d) 25 ºC at ethanol concentrations <strong>of</strong><br />

(⎯) 20, (⎯) 50, <strong>and</strong> (⎯) 80% (v/v), respectively. Figure 2e shows the variation <strong>of</strong> the hydrodynamic radii <strong>of</strong> the main peak <strong>of</strong> the<br />

distribution with ethanol concentration after incubation at 65º C <strong>and</strong> () pH 7.4 <strong>and</strong> () 2.0, or at 25 ºC <strong>and</strong> () pH 7.4 <strong>and</strong> () 2.0.<br />

are bare size estimations. In general, aggregates formed at<br />

physiological pH are bulkier if compared to those formed under<br />

acidic conditions, as plotted in Figure 2e. Structures with rh,app =<br />

400-1000 nm constitute the main contribution to light scattering.<br />

As depicted in the TEM images shown below, vide infra, these<br />

can be associated with fibrillar conformations, in particular,<br />

assigned to the large axis <strong>of</strong> protein fibers. Those values <strong>of</strong> rh,app between 30-50 nm might have different interpretations depending<br />

on solution conditions: On one h<strong>and</strong>, at acidic pH <strong>and</strong> 25 ºC they<br />

correspond to oligomeric structures, which confirms the two-step<br />

20<br />

25<br />

aggregation process observed from SLS data under these<br />

conciditions; on the other, at high temperature conditions rh,app values might also include contributions from the fibrillar<br />

aggregates width. Finally, the smallest rh,app values around 2-5 nm<br />

fit fairly well with the estimated size <strong>of</strong> protein monomers.<br />

Conformational structure changes along protein fibrillation.<br />

To get further insight into the structural characterization <strong>of</strong> the<br />

resulting protein aggregates, CD, FT-IR <strong>and</strong> ThT fluorescence<br />

d)<br />

b)


5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

55<br />

Figure 3: Selected far-UV spectra <strong>of</strong> HSA solutions after incubation at pH 7.4 at a) 65 ºC <strong>and</strong> b) 25 ºC; <strong>and</strong> at pH 2.0 at c) 65 ºC <strong>and</strong> d)<br />

25 ºC at different ethanol concentrations in the mixed solvent: 1) 20, 2) 40, 3) 60, <strong>and</strong> 4) 80% (v/v).<br />

experiments were performed. Figure 3 represents the far-UV CD<br />

spectra <strong>of</strong> HSA solutions at the different experimental conditions.<br />

When incubation is performed at high temperature <strong>and</strong><br />

physiological pH (Figure 3a), a minimum at ca. 215 nm is<br />

observed in the whole range <strong>of</strong> solvent mixture compositions.<br />

This minimum denotes the predominant presence <strong>of</strong> β-sheet<br />

structure. The amount <strong>of</strong> α-helix conformation remains constant<br />

up to a 50% (v/v) alcohol concentration (see Table S2) <strong>and</strong>, then,<br />

it slightly decreases with an additional slight gain <strong>of</strong> β-sheets. At<br />

the largest ethanol concentrations (> 80% v/v), the reduction<br />

observed in ellipticity is related to light scattering caused by the<br />

presence <strong>of</strong> big aggregates in solution. These aggregates do not<br />

enable the obtention <strong>of</strong> reliable structural composition values.<br />

Upon incubation at pH 7.4 <strong>and</strong> 25 ºC (Figure 3b), a little increase<br />

in α-helical content is observed at ethanol concentrations below<br />

30% (v/v) as a result <strong>of</strong> the weakening <strong>of</strong> hydrophobic<br />

interactions due to the presence <strong>of</strong> the cosolvent. 36 This fact<br />

results in the stabilization <strong>of</strong> the compact protein native structure<br />

<strong>and</strong> strengthens other interactions such as hydrogen bonding,<br />

stabilizing the secondary structure. Thus, alcohol acts as a<br />

“structure maker”. At larger ethanol concentrations (30-60 %<br />

(v/v)) a simultaneous decrease in α-helix <strong>and</strong> an increase in β-<br />

sheet structures take place in a wider alcohol range as compared<br />

to non-buffered solutions 36 , which is characterized by the absence<br />

<strong>of</strong> the minimum at 220-224 nm <strong>and</strong> the appearance <strong>of</strong> a new one<br />

at ca. 215 nm 60,61 (see Figure 3b). At ethanol contents higher than<br />

60% (v/v), a progressive new increase in ellipticity at 222 nm <strong>and</strong><br />

208 nm occurs, which suggests the formation <strong>of</strong> non-native<br />

helical structure at the expense <strong>of</strong> β-sheet <strong>and</strong> unordered<br />

conformations. The loss <strong>of</strong> non-polar contacts at high ethanol<br />

concentrations appears to favor the formation <strong>of</strong> intramolecular<br />

hydrogen bonds responsible <strong>of</strong> the formation <strong>of</strong> new helices. This<br />

non-native α-helix structure is more prone to aggregation, 62 as<br />

denoted by the important increase in aggregate sizes previously<br />

observed in Figure 2.<br />

Incubation at acidic pH <strong>and</strong> high temperature (Figure 3c) was<br />

found to involve an increasingly negative Cotton effect at 208 nm<br />

<strong>and</strong> the simultaneous decay in ellipticity at 222 nm. This suggests<br />

that, while ethanol induces disorder, some helical content is still<br />

retained. In fact, within the ethanol concentration range between<br />

20-50% (v/v), the α-helix <strong>and</strong> unordered conformation contents<br />

slightly increase at the expense <strong>of</strong> β-sheets (see Table S2). The<br />

formation <strong>of</strong> this extra α-helix might arise from the disruption <strong>of</strong><br />

long-range hydrophobic interactions due to the presence <strong>of</strong><br />

alcohol molecules, which allows local helical conformation to<br />

provide a transient means <strong>of</strong> decreasing backbone solvation. At<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 7<br />

35<br />

40<br />

45<br />

50<br />

191


5<br />

10<br />

15<br />

20<br />

25<br />

Figure 4: Normalized maximum ThT fluorescence intensities <strong>of</strong> HSA samples incubated at different ethanol concentrations in the mixed<br />

solvent at a) pH 7.4 <strong>and</strong> 65 ºC, b) pH 7.4 <strong>and</strong> 25 ºC, c) pH 2.0 <strong>and</strong> 65 ºC <strong>and</strong> d) pH 2.0 <strong>and</strong> 25 ºC. Normalization is made with respect to<br />

the maximum ThT fluorescence observed at pH 7.4 <strong>and</strong> 65 ºC.<br />

30 ethanol concentrations larger than 50% (v/v) the β-sheet content<br />

increases, as denoted from the progressive shift <strong>of</strong> the minimum<br />

in CD spectra to 215 nm, which is characteristic <strong>of</strong> β-str<strong>and</strong><br />

formation. In contrast, at acidic pH <strong>and</strong> 25 ºC (Figure 3d), the<br />

structural content <strong>of</strong> protein samples at ethanol concentrations<br />

35 lower than 60% (v/v) remains almost invariable. Since HSA<br />

molecules are in a starting acid-denaturated state, alcohol<br />

stabilizes the α-helix conformation <strong>of</strong> the acid-unfolded protein<br />

monomers by minimizing the exposure <strong>of</strong> the peptide backbone.<br />

In particular, at alcohol concentrations < 40% (v/v),<br />

40 intermolecular interactions are disfavored by suppression <strong>of</strong> the<br />

strong aggregate-stabilizing effect <strong>of</strong> negatively charged residues,<br />

still resulting in an effective electrostatic repulsion between<br />

positively charged chains. Hence, under these conditions alcohol<br />

stabilizes the molten-globule state <strong>of</strong> HSA during incubation, <strong>and</strong><br />

only protein clusters can be formed. 63,64 45<br />

The presence <strong>of</strong> these<br />

clusters, confirmed by the presence <strong>of</strong> a small peak at relatively<br />

low aggregate sizes (∼20-30 nm) (see Figures 2d), implies some<br />

increase in light scattering from solution as shown previously<br />

<strong>and</strong>, thus, possibly originates the slight decrease in ellipticity<br />

50 observed in Figure 3d. Protein molecules experience additional<br />

structural rearrangements at ethanol concentrations between 50-<br />

90% (v/v) when the polarity <strong>of</strong> the medium is drastically<br />

changed. This involves a decrease in α-helix structure <strong>and</strong><br />

formation <strong>of</strong> β-str<strong>and</strong>s that, as observed from light scattering<br />

55 data, are prone to aggregate. FT-IR measurements corroborate the<br />

structural changes undergone by protein molecules depicted by<br />

CD data (for additional comments see ESI).<br />

Although reliable in detecting β-str<strong>and</strong>s <strong>and</strong> probing hydrogen<br />

bonding between them, CD <strong>and</strong> FT-IR spectroscopies fail to<br />

60 discriminate between amorphous <strong>and</strong> ordered aggregates. Since<br />

ThT dye strongly emits when bound to amyloid-like material<br />

rather than to amorphous aggregates, we conducted ThT<br />

fluorescence measurements <strong>of</strong> HSA samples under the different<br />

solution conditions in order to determine the presence <strong>of</strong> ordered<br />

65 aggregates (amyloid-like fibrils) in solution. Experimental data<br />

were background corrected for solvent <strong>and</strong> monomeric protein<br />

contributions <strong>and</strong> normalized to the largest value attained, i.e., at<br />

pH 7.4 <strong>and</strong> 65 ºC at 70 % (v/v) ethanol. Figure 4 confirms that βsheet<br />

rich structures (fibrils as shown by TEM pictures below) are<br />

70 formed under all conditions except at acidic pH <strong>and</strong> 25 ºC at<br />

alcohol concentrations below 60% (v/v). At pH 7.4 <strong>and</strong> 65 ºC<br />

(Figure 4a), the amount <strong>of</strong> fibrils progressively increases up to an<br />

ethanol content <strong>of</strong> 70% (v/v), <strong>and</strong> then decreases. This decrease<br />

in ThT fluorescence can be associated with the presence <strong>of</strong><br />

75 mature fibril association/aggregation, as confirmed by TEM (see<br />

below), which may reduce the protein exposed surface area <strong>and</strong><br />

the number <strong>of</strong> available ThT binding sites. 65 In contrast, at acidic<br />

conditions (Figure 4c) the maximum level <strong>of</strong> fibril formation<br />

takes place at the largest ethanol concentrations (70-90% v/v). On<br />

80 the other h<strong>and</strong>, the temporal evolution <strong>of</strong> ThT binding confirms<br />

8 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society <strong>of</strong> Chemistry [year]<br />

192


10<br />

15<br />

S<strong>of</strong>t Matter<br />

Cite this: DOI: 10.1039/c0xx00000x<br />

www.rsc.org/xxxxxx<br />

a)<br />

c) d)<br />

e) f)<br />

Dynamic Article Links ►<br />

PAPER<br />

Figure 5: Selected TEM images <strong>of</strong> HSA samples incubated at pH 7.4 <strong>and</strong> a-c) 65 ºC, or d-f) 25 ºC at ethanol concentrations in the mixed<br />

solvent <strong>of</strong> a,d) 20, b,e) 60, c,f) 80% (v/v), respectively.<br />

the trends previously observed from SLS kinetic data, although<br />

some differences at acidic <strong>and</strong> high temperature conditions are<br />

observed (see Figure S2 <strong>and</strong> further explanation in ESI).<br />

Based on all the above findings, the combination <strong>of</strong> i)<br />

incubation at high temperature <strong>and</strong> ii) the presence <strong>of</strong> ethanol as a<br />

cosolvent in solution result in important variations <strong>of</strong> the protein<br />

structural content <strong>and</strong> subsequent strengthening <strong>of</strong> hydrophobic<br />

interactions between protein molecules. Incubation at high<br />

temperatures involves the denaturation <strong>of</strong> protein monomers due<br />

to the disruption <strong>of</strong> the structure <strong>of</strong> protein domains 48,49,63 while<br />

the presence <strong>of</strong> the cosolvent changes the solvent polarity <strong>and</strong><br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 9<br />

193<br />

5<br />

20<br />

25<br />

30<br />

b)<br />

induces a further exposure <strong>of</strong> residues, 66 which enhances the<br />

internal degrees <strong>of</strong> freedom <strong>of</strong> side-chains. 67 Then, in most <strong>of</strong> the<br />

studied conditions, the cosolvent induces additional modifications<br />

in the starting protein structure, leading to a strong protein<br />

aggregation <strong>and</strong> fibril formation, but with differences in the<br />

packing mechanisms leading to fibril polymorphism, as will be<br />

shown below. All these facts become even clearer when<br />

comparing the secondary structure compositions here obtained<br />

with those measured under identical incubation conditions in the<br />

absence <strong>of</strong> alcohol. 36,40-42


Aggregate morphology.<br />

In order to determine the structural arrangement <strong>of</strong> protein<br />

aggregates <strong>and</strong> confirm the presence <strong>of</strong> amyloid-like fibrils, TEM<br />

5 images were obtained. Figures 5 <strong>and</strong> 6 show the structures <strong>of</strong> the<br />

different protein aggregates upon incubation under the different<br />

mixed solvent conditions.<br />

Samples incubated at 65 ºC <strong>and</strong> physiological pH fibrillated<br />

in the whole range <strong>of</strong> alcohol concentrations analyzed (Figure 5a-<br />

10 c). However, remarkable differences in the structure <strong>of</strong> the<br />

resulting fibrils are observed if compared to those obtained in the<br />

absence <strong>of</strong> ethanol. In particular, for alcohol concentrations up to<br />

70% (v/v) straight-several micrometers-long fibers with widths <strong>of</strong><br />

ca. 10-12 nm are seen (Figure 5a,b). These fibrils are longer than<br />

15 those obtained in the absence <strong>of</strong> ethanol under similar conditions<br />

<strong>and</strong> do not possess the typically observed curly morphology <strong>of</strong><br />

HSA fibrils; 40,68 In fact, they resemble those fibers obtained for<br />

classically amyloidogenic proteins such as α-synuclein, 69<br />

insulin, 67 or lysozyme. 70 Probably, the larger β-sheet content<br />

20 which results from the combination <strong>of</strong> high temperature <strong>and</strong><br />

alcohol-mixed solvent conditions during incubation favored the<br />

establishment <strong>of</strong> strong attractive interactions between protein<br />

monomers <strong>and</strong> a further exposition <strong>of</strong> their side chains to solvent.<br />

The reduction <strong>of</strong> the thermodynamic costs for the exposure <strong>of</strong><br />

25 hydrophobic residues as the dielectric properties <strong>of</strong> the solvent<br />

approach to those <strong>of</strong> the bulk alcohol may be the origin for such<br />

behavior. Hence, this enables the possibility <strong>of</strong> a better packing <strong>of</strong><br />

the protein str<strong>and</strong>s inside the fibrils favoring an increase in fibril<br />

length. At larger ethanol contents (> 70% v/v), an important<br />

30 amount <strong>of</strong> thicker mature fibrils formed by lateral assembly (side<br />

by side) <strong>of</strong> two or more individual fibrils is observed. Moreover,<br />

very large aggregates composed <strong>of</strong> assembled fibrils are also seen<br />

(Figure 5c), which agrees with the decrease in ThT fluorescence<br />

<strong>and</strong> the existence <strong>of</strong> a third population distribution at very large<br />

35 sizes observed by DLS.<br />

As observed from Figure 5d-f, incubation at room<br />

temperature <strong>and</strong> physiological pH involved the formation <strong>of</strong><br />

protein globules (with sizes ranging from 15 to 60 nm), which<br />

seem to elongate at low alcohol content (< 30% v/v) as a<br />

40 consequence <strong>of</strong> the existence <strong>of</strong> attractive interactions between<br />

them (see Figure 5d). As ethanol concentration increases (up to<br />

60% v/v) we observe the formation <strong>of</strong> curly fibrils with similar<br />

characteristic features (lengths between 0.2-1 µm, widths ranging<br />

between 8-15 nm, <strong>and</strong> a curly morphology) as those obtained<br />

45 upon incubation at high temperature in the absence <strong>of</strong> the alcohol<br />

(Figure 5e). 40 At larger ethanol contents (>60% v/v) fibrils<br />

progressively disappeared <strong>and</strong> large-elongated glomerular<br />

aggregates were observed. Considering CD, FTIR, <strong>and</strong> ThT<br />

fluorescence data obtained at this alcohol concentration range, we<br />

50 expect these aggregates to possess an increased level <strong>of</strong> α-helix<br />

content (Figure 5f) <strong>and</strong> recover the tertiary conformation. 36<br />

Under acidic conditions <strong>and</strong> 65 ºC, more curly fibrils are<br />

observed if compared with incubation at 25 ºC (see Figure 6a <strong>and</strong><br />

d, for example). Also, the width <strong>and</strong> length <strong>of</strong> these fibrils are<br />

55 slightly larger than at room temperature. At intermediate ethanol<br />

contents, interconnected networks <strong>of</strong> elongated aggregates are<br />

observed (Figure 6b), which agrees with the increases in scattered<br />

light observed by SLS data. At large ethanol concentrations (><br />

60% v/v), single dispersed curly fibrils are again observed<br />

60 (Figure 6c), in agreement with the decrease in SLS intensities <strong>and</strong><br />

conformational CD <strong>and</strong> FTIR data. On the other h<strong>and</strong>, at acidic<br />

pH <strong>and</strong> 25 ºC oligomeric partially-elongated structures with<br />

lengths <strong>of</strong> ca. 20-40 nm <strong>and</strong> widths <strong>of</strong> 3-4 nm are observed at<br />

ethanol concentrations between 0-40% (v/v), as seen in the<br />

absence <strong>of</strong> alcohol (see Figure 6d). 40 According to ThT <strong>and</strong> CD<br />

data, these structures are mainly composed <strong>of</strong> α-helix <strong>and</strong><br />

unordered conformations. When larger amounts <strong>of</strong> ethanol are<br />

present in the solvent (>60% v/v), fibrils with a curly morphology<br />

are present (see Figures 6e,f). Under the present conditions, it<br />

seems the increasing amounts <strong>of</strong> ethanol induce an enlargement<br />

<strong>and</strong> fusion <strong>of</strong> oligomers as a result <strong>of</strong> additional intermolecular<br />

hydrophobic contacts between side chains, which favors the<br />

formation <strong>of</strong> β-str<strong>and</strong>s <strong>and</strong>, hence, fibril formation.<br />

10 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society <strong>of</strong> Chemistry [year]<br />

65<br />

70<br />

75<br />

80<br />

85<br />

90<br />

95<br />

100<br />

105<br />

110<br />

115<br />

Conclusions<br />

In summary, from a general perspective, ordered<br />

aggregation <strong>of</strong> globular proteins requires partial unfolding <strong>of</strong> the<br />

native state. 71 This enables the globular precursor state to be<br />

populated <strong>and</strong> to expose aggregation-competent regions that are<br />

usually protected against intermolecular interactions in the native<br />

protein. However, these regions can be different <strong>and</strong> display<br />

distinct propensities to solvent exposition depending on solution<br />

conditions <strong>and</strong>, thus, different ability to fibrillate. 40,72 This seems<br />

to be the origin <strong>of</strong> the differences observed in the behavior <strong>of</strong><br />

HSA samples incubated at physiological <strong>and</strong> acidic pH at the<br />

different solvent compositions. Under physiological pH ethanol is<br />

able to destabilize the protein secondary structure in most <strong>of</strong> the<br />

conditions, giving rise to an aggregation/fibrillation process. In<br />

contrast, under acidic conditions pH induces further structural<br />

changes in secondary structure favoring electrostatic repulsive<br />

interactions, which makes the protein molecules more soluble,<br />

disfavoring their aggregation to a great extent. Only when a<br />

cosolvent concentration threshold is reached does the fibrillation<br />

process start to be significant. Despite amyloid formation can<br />

occur under different solution conditions, rates <strong>and</strong> extents <strong>of</strong><br />

fibril formation, fibrillation pathways, <strong>and</strong> the structures <strong>of</strong><br />

resulting fibrils can be very diverse depending on solution<br />

conditions. Constraints due to electrostatic <strong>and</strong> hydrophobic<br />

interactions are, in this regard, important in that they govern the<br />

side-chain packing that controls stacking <strong>of</strong> β-sheets, favoring the<br />

formation <strong>of</strong> more classical amyloid-like fibrils at physiological<br />

pH <strong>and</strong> high temperature. The different modes <strong>of</strong> association<br />

under the present conditions are also dictated by the influence <strong>of</strong><br />

solvent on hydrogen bonding within the β-sheets. At this respect,<br />

the distinct self-assembled motifs can be also a result <strong>of</strong> changes<br />

in the arrangement <strong>of</strong> β-str<strong>and</strong>s within the β-sheets. This is clear<br />

from the different types <strong>of</strong> fibrillar structures observed under<br />

varying ethanol concentrations, in particular at physiological pH:<br />

from the classical curly geometry <strong>of</strong> HSA fibrils at low ethanol<br />

concentrations to the formation <strong>of</strong> long-straight fibrils which<br />

resemble those formed in originally disease-related proteins.<br />

Nevertheless, high temperature <strong>and</strong> physiological pH conditions<br />

are the main parameters influencing the protein structure<br />

destabilization <strong>and</strong> subsequent aggregation upon incubation,<br />

while ethanol aids to regulate hydrogen bonding, the attractive<br />

hydrophobic interactions, <strong>and</strong> the protein accessible surface area.<br />

194


5<br />

10<br />

15<br />

S<strong>of</strong>t Matter<br />

Cite this: DOI: 10.1039/c0xx00000x<br />

www.rsc.org/xxxxxx<br />

a)<br />

c) d)<br />

e) f)<br />

Dynamic Article Links ►<br />

PAPER<br />

Figure 6: Selected TEM images <strong>of</strong> HSA samples incubated at pH 2.0 <strong>and</strong> a-c) 65 ºC, or d-f) at 25 ºC at ethanol concentrations in the<br />

mixed solvent <strong>of</strong> a,d) 20, b,e) 50, c,f) 80% (v/v), respectively.<br />

The role <strong>of</strong> “fibril enhancer” <strong>of</strong> ethanol under most <strong>of</strong> the<br />

conditions is contrary to what is observed when already-formed<br />

fibrils are incubated in its presence. For example, it has been<br />

shown that insulin <strong>and</strong> β-lactoglobulin fibrils disassemble upon<br />

incubation in DMSO <strong>and</strong> ethanol, respectively, as a result <strong>of</strong> a<br />

weakening <strong>of</strong> hydrogen bonding inside the fibrils. 73,74 consequence <strong>of</strong> changes in the energy l<strong>and</strong>scape <strong>of</strong> the<br />

aggregation process.<br />

20 Acknowledgement. Authors thank Ministerio de Ciencia e Innovación<br />

(MICINN) for research project MAT 2010-17336, Xunta de Galicia for<br />

research project INCITE09206020PR <strong>and</strong> for European Regional<br />

Thus, it Development Funds (research project 2010/50), <strong>and</strong> Fundación Ramón<br />

appears that the solvent composition <strong>and</strong> the different initial state Areces for additional financial support.<br />

<strong>of</strong> the protein are responsible <strong>of</strong> the observed differences as a 25<br />

Electronic Supplementary Information (ESI) available:<br />

Complementary seeding growth <strong>and</strong> concentration-dependent kinetic<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 11<br />

195<br />

b)


5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

45<br />

50<br />

55<br />

60<br />

65<br />

studies, tables containing kinetic constants derived from SLS experiments<br />

<strong>and</strong> secondary structure compositions under the different solution<br />

conditions, FTIR spectra <strong>and</strong> discussion, <strong>and</strong> temporal evolution <strong>of</strong> ThT<br />

biding with subsequent discussion. See DOI: 10.1039/b000000x/<br />

References<br />

1. T. P. Knowles, A. W. Fitzpatrick, S. Meehan, H. R. Mott, M.<br />

Vendruscolo, C. M. Dobson, M. E. Well<strong>and</strong>, Science 2007,<br />

318, 1900-1903.<br />

2. S. G. Zhang, Nature Biotechnol. 2003, 21, 1171-1178.<br />

3. A. T. Alex<strong>and</strong>rescu, Protein Sci. 2005, 14, 1-12.<br />

4. K. Lundmark, G. T. Westermark, A. Olsen, P. Westermark,<br />

Proc. Natl. Acad. Sci. USA 2005, 102, 6098-6102.<br />

5. M. B. Pepys, Ann. Rev. Med. 2006, 57, 223-241.<br />

6. M. Calamai, F. Chiti, C. M. Dobson, Biophys. J. 2005, 89,<br />

4201-4210.<br />

7. B. A. Vernaglia, J. Huang, E. D. Clark, Biomacromolecules<br />

2004, 5, 1362-1370.<br />

8. S. Goda, K. Takano, K. Yutani, Y. Yamagata, R. Nagata, H.<br />

Akutsu, S. Maki, K. Namba, Protein Sci. 2000, 9, 369-375.<br />

9. F. G. de Felice, M. N. N. Vieira, M. N. L. Meirelles, L. A.<br />

Morozova-Roche, C. M. Dobson, S. T. Ferreira, FASEB J.<br />

2004, 18, 1099-1101.<br />

10. A. E. Cao, D. Y. Hu, L. H. Lai, Protein Sci. 2004, 13, 319-324.<br />

11. M. R. H. Krebs, D. K. Wilkins, E. W. Chung, M. C.<br />

Pitkeathly, A. K. Chamberlain, J. Zurdo, C. V. Robinson, C.<br />

M. Dobson, J. Mol. Biol. 2000, 300, 541-549.<br />

12. F. Bemporad, G. Calloni, S. Campioni, G. Plakoutsi, N.<br />

Taddei, F. Chiti, Acc. Chem. Res. 2006, 39, 620-627.<br />

13. G. Soldi, F. Bemporad, S. Torrassa, A. Relini, M. Ramazzotti,<br />

N. Taddei, F. Chiti, Biophys. J. 2005, 89, 4234-4244.<br />

14. S. N. Timasheff, Ann. Rev. Biophys. Biomol. Struct. 1993, 22,<br />

67-97.<br />

15. W. Dzwolak, Biochim. Biophys. Acta 2006, 1764, 470-480.<br />

16. N. Rezaei-Ghaleh, M. Zweckstetter, D. Morshedi, A. Ebrahim-<br />

Habibi, M. Nemat-Gorgani, Biopolymers 2009, 91, 28-36.<br />

17. J. H. Lee, G. Bhak, S. G. Lee, S. R. Paik, Biophys. J. 2008, 95,<br />

L16-L18.<br />

18. M. Holley, C. Eginton, D. Schaefer, L. R. Brown, Biochem.<br />

Biophys. Res. Comm. 2008, 373, 164-168.<br />

19. V. Castelletto, I. W. Hamley, P. J. F. Harris, U. Olsson, N. J.<br />

Spencer, Phys. Chem. B 2009, 113, 9978-9987.<br />

20. I. W. Hamley, D. R. Nutt, G. D. Brown, J. F. Miravet, B.<br />

Escuder, F. Rodríguez-Llansola, J. Phys. Chem. B 2009, 114,<br />

940-951.<br />

21. A. K. Paravastu, A. T. Petkova, R. Tycko, Biophys. J. 2006,<br />

90, 4618-4629.<br />

22. W. Hoyer, T. Antony, D. Cherny, G. Heim, T. M. Jovin, V.<br />

Subramaniam, J. Mol. Biol. 2002, 322, 383-393.<br />

23. D. P. Smith, S. Jones, L. C. Serpell, M. Sunde, S. E. Radford,<br />

J. Mol. Biol. 2003, 330, 943-954.<br />

24. J. S. Patton, R. M. Platz, Adv. Drug Delivery Rev.1992, 8, 179-<br />

196.<br />

25. W. S. Choi, G. G. K. Murthy, D. A. Edwards, R. Langer, A.<br />

M. Klibanov, Proc. Natl. Acad. Sci. USA 2001, 98, 11103-<br />

11107.<br />

26. S. Cinelli, G. Onori, A. Santucci, Colloids Surfaces A 1999,<br />

160, 3-8.<br />

27. V. E. Bychkova, O. B. Ptitsyn, Biochem. Mol. Biol. 1993, 4,<br />

133-163<br />

28. O. B. Ptitsyn, V. E. Bychkova, V. N. Uversky, Philos. Trans.<br />

Royal Soc. London B-Biol. Sci. 1995, 348, 35-41.<br />

29. V. E. Bychkova, A. E. Dujsekina, S. I. Klenin, E. E. Tiktopulo,<br />

V. N. Uversky, O. B. Ptitsyn, Biochemistry 1996, 35, 19,<br />

6058-6063.<br />

30. V. N. Uversky, N. V. Narizhneva, S. O. Kirschstein, S. Winter,<br />

G. Löber, Folding Design 1997, 2, 163-172.<br />

31. Y. O. Kamatari, T. Konno, M. Kataoka, K. Akasaka, J. Mol.<br />

70 Biol. 1996, 259, 512-523.<br />

32. N. V. Narizhneva, V. N. Uversky, Protein Pept. Lett. 1997, 4,<br />

243-249.<br />

33. Y. X. Qu, C. L. Bolen, D. W. Bolen, Proc. Natl. Acad. Sci.<br />

USA 1998, 95, 9268-9273.<br />

34. S. N. Timasheff, Proc. Natl. Acad. Sci. USA 2002, 99, 9721-<br />

12 | Journal Name, [year], [vol], 00–00 This journal is © The Royal Society <strong>of</strong> Chemistry [year]<br />

75<br />

80<br />

85<br />

90<br />

95<br />

100<br />

105<br />

110<br />

115<br />

120<br />

125<br />

130<br />

135<br />

9726.<br />

35. D. Liu, T. Wyttenbach, C. J. Carpenter, M. T. Bowers, J. Am.<br />

Chem. Soc. 2004, 126, 3261-3270.<br />

36. P. Taboada, S. Barbosa, E. Castro, M. Gutiérrez-Pichel, V.<br />

Mosquera, Chem. Phys. 2007, 340, 59-68.<br />

37. P. Taboada, S. Barbosa, E. Castro, V. Mosquera, J. Phys.<br />

Chem. B 2006, 110, 20733-20736.<br />

38. Y. Aso, K. Shiraki, M. Takagi, Biosci. Biotech. Biochem.<br />

2007, 71, 1313-1321.<br />

39. W. Dzwolak, S. Grudzielanek, V. Smirnovas, R. Ravindra, C.<br />

Nicolini, R. Jansen, A. Loksztejn, S. Porowski, R. Winter,<br />

Biochemistry 2005, 44, 8948-8958.<br />

40. J. Juárez, P. Taboada, V. Mosquera, Biophys. J. 2009, 96,<br />

2353-2370.<br />

41. J. Juárez, S. Goy-López, A. Cambón, P. Taboada, V.<br />

Mosquera, J. Phys. Chem. B 2009, 113, 10521-10529.<br />

42. J. Juárez, P. Taboada, S. Goy-López, A. Cambón, M.-B.<br />

Madec, S. G. Yeates, V. Mosquera, J. Phys. Chem. B 2009,<br />

113, 12391-12399.<br />

43. C. N. Pace, F. Vajdos, L. Fee, G. Grimsley, T. Gray, Protein<br />

Sci. 1995, 4, 2411-2423.<br />

44. S. W. Provencher, Comp. Phys. Comm. 1982, 27, 213-227.<br />

45. A. Lobley, L. Whitmore, B. A. Wallace, Bioinformatics 2002,<br />

18, 211-212.<br />

46. L. Whitmore, B. A. Wallace, Nucleic Acids Res. 2004, 32,<br />

W668-W673.<br />

47. M. J. Geisow, G. H. Beaven, Biochem. J. 1977, 163, 477-484.<br />

48. B. Farruggia, F. Rodriguez, R. Rigatuso, G. Fidelio, G. Pico, J.<br />

Protein Chem. 2001, 20, 81-89.<br />

49. K. Flora, J. D. Brennan, G. A. Baker, M. A. Doody, F. V.<br />

Bright, Biophys. J. 1998, 75, 1084-1096.<br />

50. N. K. Holm, S. K. Jespersen, L. V. Thomassen, T. Y. Wolff, P.<br />

Sehgal, L. A. Thomsen, G. Christiansen, C. B. Andersen, A.<br />

D. Knudsen, D. E. Otzen, Biochim. Biophys. Acta 2007, 1774,<br />

1128-1138.<br />

51. F. Ferrone, W. Ronald, Analysis <strong>of</strong> protein aggregation<br />

kinetics. In Methods in Enzymology, Academic Press: 1999,<br />

309, 256-274.<br />

52. M. Alatorre-Meda, P. Taboada, J. Sabin, B. Krajewska, L. M.<br />

Varela, J. R. Rodriguez, Colloids Surfaces A 2009, 339, 145-<br />

152.<br />

53. P. Shahi, R. Sharma, S. Sanger, I. Kumar, R. S. Jolly,<br />

Biochemistry 2007, 46, 7365-7373.<br />

54. J. T. Jarrett, P. T. Lansbury, Biochemistry 1992, 31, 12345-<br />

12352.<br />

55. A. Lomakin, D. B. Teplow, D. A. Kirschner, G. B. Benedek,<br />

Proc. Natl. Acad. Sci. USA 1997, 94, 7942-7947.<br />

56. R. Bader, R. Bamford, J. Zurdo, B. F. Luisi, C. M. Dobson, J.<br />

Mol. Biol. 2006, 356, 189-208.<br />

57. Q. Xu, T. A. Keiderling, Biochemistry 2005, 44, 7976-7987.<br />

58. A. J. Rader, B. M. Hespenheide, L. A. Kuhn, M. F. Thorpe,<br />

Proc. Natl. Acad. Sci. USA 2002, 99, 3540-3545.<br />

59. L. Nielsen, R. Khurana, A. Coats, S. Frokjaer, J. Brange, S.<br />

Vyas, V. N. Uversky, A. L. Fink, Biochemistry 2001, 40,<br />

6036-6046.<br />

60. S. Goda, K. Takano, Y. Yamagata, R. Nagata, H. Akutsu, S.<br />

Maki, K. Namba, K. Yutani, Protein Sci. 2000, 9, 369-375.<br />

61. Y. Yonezawa, S. Tanaka, T. Kubota, K. Wakabayashi, K.<br />

Yutani, S. Fujiwara, J. Mol. Biol. 2002, 323, 237-251.<br />

62. N. A. P. Ulrih, G. Anderluh, P. Macek, T. V. Chalikian,<br />

Biochemistry 2004, 43, 9536-9545.<br />

63. M. Dockal, D. C. Carter, F. Ruker, J. Biol. Chem. 2000, 275,<br />

3042-3050.<br />

64. Y. Kumar, S. Tayyab, S. Muzammil, Archiv. Biochem.<br />

140 Biophys. 2004, 426, 3-10.<br />

196


5<br />

65. L. A. Sikkink, M. Ramirez-Alvarado, Biophys. Chem. 2008,<br />

135, 25-31.<br />

66. I. I. Stepuro, E. A. Lapshina, N. A. Chaikovskaia, Mol Biol<br />

(Mosk) 1991, 25, 337-347.<br />

67. S. Grudzielanek, R. Jansen, R. Winter, J. Mol. Biol. 2005, 351,<br />

879-894.<br />

68. L. M. C. Sagis, C. Veerman, E. van der Linden, Langmuir<br />

2004, 20, 924-927.<br />

69. Y. Fezoui, D. B. Teplow, J. Biol. Chem. 2002, 277, 36948-<br />

10 36954.<br />

25<br />

70. K. Sasahara, H. Yagi, H. Naiki, Y. Goto, Y. J. Mol. Biol. 2007,<br />

372, 981-991.<br />

71. V. N. Uversky, A. L. Fink, Biochim. Biophys. Acta 2004,<br />

1698, 131-153.<br />

72. M. Bhattacharya, N. Jain, S. Mukhopadhyay, J. Phys. Chem. B<br />

This journal is © The Royal Society <strong>of</strong> Chemistry [year] Journal Name, [year], [vol], 00–00 | 13<br />

15<br />

2011, 115, 4195-4205.<br />

73. N. Hirota-Nakaoka, K. Hasegawa, H. Naiki, Y. Goto, J.<br />

Biochem. 2003, 134, 159-164.<br />

74. S. Jordens, J. Adamcik, I. Amar-Yuli, R. Mezzenga,<br />

20 Biomacromolecules 2011, 12, 187-193.<br />

197


Obtention <strong>of</strong> Metallic Nanowires by Protein<br />

Biotemplating <strong>and</strong> Their Catalytic Application<br />

Josue Juarez, Adriana Cambon, Sonia Goy-Lopez, Antonio Topete,<br />

Pablo Taboada,* <strong>and</strong> Víctor Mosquera<br />

Grupo de Física de Coloides y Polímeros, Departamento de Física de la Materia Condensada, Facultad de Física,<br />

Universidad de Santiago de Compostela, E-15782 Santiago de Compostela, Spain<br />

ABSTRACT Gold nanowires were obtained by a seeded growth process using<br />

protein lysozyme fibrils as biotemplates. The degree <strong>of</strong> metal coverage until full<br />

biotemplate coverage onto the fibril was controlled by the gold salt concentration<br />

<strong>and</strong> the number <strong>of</strong> sequential additions <strong>of</strong> metal growth solutions. The hybrid<br />

fibrils might have a potential use as catalysts since they display enhanced catalytic<br />

activity in the reduction <strong>of</strong> p-nitrophenol to p-aminophenol by NaBH 4.<br />

SECTION Nanoparticles <strong>and</strong> Nanostructures<br />

Metal nanoparticles have received considerable attention<br />

in past decade because <strong>of</strong> their particular optical,<br />

electronic, magnetic, <strong>and</strong> catalytic properties <strong>and</strong> their<br />

important applications in many fields such as nanosensors, 1-4<br />

catalysis, 5-8 biomedicine, 9-11 biological labeling, 12-14 <strong>and</strong><br />

surface-enhanced Raman scattering (SERS). 15-18 To optimize<br />

<strong>and</strong> extend the application <strong>of</strong> metal NPs, methods must be<br />

developed to control the assembly <strong>and</strong> organization <strong>of</strong> these<br />

nanomaterials. Assemblies <strong>of</strong> NPs provide optical <strong>and</strong> electronic<br />

properties that are distinct compared to individual<br />

particles or disorganized macroscale agglomeration. In this<br />

regard, Nature <strong>of</strong>fers us the possibility to take advantage <strong>of</strong> the<br />

well-defined structures <strong>and</strong> special properties <strong>of</strong> biomolecules<br />

<strong>and</strong> their supramolecular structures to organize nanoparticles<br />

into predefined, topologically intricate nanostructures or to<br />

synthesize miscellaneous materials in order to control the<br />

properties <strong>of</strong> nanoparticle assemblies for potential applications<br />

in electronic, optical, <strong>and</strong> chemical devices. 19-23 In<br />

particular, functionalizing one-dimensional (1D) supporting<br />

biomaterials with metal NPs that combine the properties <strong>of</strong><br />

two functional nanomaterials, such as high conductivity surface<br />

area or precise chemical functionality <strong>of</strong> the biotemplate<br />

<strong>and</strong> the unique plasmonic or catalytic properties <strong>of</strong> metal NPs,<br />

to achieve a wider range <strong>of</strong> applications will therefore play an<br />

important role in the development <strong>of</strong> nanoscience <strong>and</strong> nanotechnology.<br />

As a result, considerable efforts have been directed<br />

toward the use <strong>of</strong> 1D biotemplates such as DNA, 24-27<br />

viruses, 28-31 peptides, 32-34 proteins, 35-38 or fungus 39 to construct<br />

corresponding multifunctional hybrid 1D nanostructures<br />

such as nanowires, 40-43 nanotubes, 44-46 or nanodumbbells. 47<br />

The spontaneous fibrillation mechanism which different<br />

proteins undergo under suitable conditions results in the<br />

formation <strong>of</strong> amyloid-like fibrils, 37,48,49 which should be valid<br />

as alternative templating agents for the construction <strong>of</strong> 1D<br />

hybrid materials. In particular, lysozyme (Lys) is a protein<br />

which is known to fibrillate under different solution conditions<br />

pubs.acs.org/JPCL<br />

to form protein fibrils with diameters ranging from 10 to 50 nm<br />

<strong>and</strong> lengths up to several micrometers (see Figure 1a). 50-52<br />

The dimensions <strong>of</strong> the fibrils can be tuned by varying the<br />

solution conditions <strong>and</strong> the incubation time. They are also<br />

thermally stable at high temperatures <strong>and</strong> different pHs, <strong>and</strong><br />

they contain multiple potential ion-binding sites within their<br />

amino acid sequences, which enables their use in metallization<br />

reactions under relatively harsh conditions. Thus, we<br />

exploit all <strong>of</strong> these features to generate 1D nanoparticle<br />

assemblies <strong>and</strong> nanowires by in situ nanoparticle formation<br />

<strong>and</strong> seeding growth on the fibril surface <strong>and</strong> to analyze their<br />

potential applicability as biocatalysts.<br />

Although biotemplating <strong>of</strong> 1D metallic nanoparticle assemblies<br />

<strong>and</strong> nanowires by peptides <strong>and</strong> protein fibrils has<br />

been previously reported 35,37,38,40,43 <strong>and</strong> their electric <strong>and</strong><br />

magnetic properties have been evaluated, 35,53 less attention<br />

has been paid to the obtention <strong>of</strong> 1D protein-fibril-templated<br />

metallic nanostructures with a controlled degree <strong>of</strong> metallization<br />

<strong>and</strong> their potential application as a catalyst. Here, we<br />

present a seeded-mediated method for the controlled coverage<br />

by Au NPs on the surface <strong>of</strong> the fibrils formed by the<br />

protein Lys in order to find the optimal conditions for the use<br />

<strong>of</strong> the present biohybrid systems as a catalyst. We test their<br />

catalytic capability using the model reduction reaction <strong>of</strong><br />

p-nitrophenol by NaBH 4.<br />

Lys fibrils were prepared as previously described (see<br />

Experimental Section). Au-nanoparticle-coated protein fibrils<br />

were obtained by the addition <strong>of</strong> preformed citrate-stabilized<br />

Au seeds to an aqueous solution containing self-assembled<br />

fibrils at pH 2.0. Provided that the isoelectric point <strong>of</strong> Lys is<br />

around 11, 54 it is expected that negatively charged gold seed<br />

Received Date: July 26, 2010<br />

Accepted Date: August 23, 2010<br />

r XXXX American Chemical Society 2680 DOI: 10.1021/jz101029f |J. Phys. Chem. Lett. 2010, 1, 2680–2687<br />

199


ions would interact electrostatically with electrically charged<br />

amine groups located at the fibril surfaces after incubation.<br />

Subsequently, different sequential additions <strong>of</strong> a solution<br />

containing HAuCl 4 <strong>and</strong> ascorbic acid were performed in order<br />

to achieve a full coverage <strong>of</strong> the hybrid composites. Different<br />

initial [Au seeds]/[protein] molar ratios (MR) were tested in<br />

order to find an initial suitable MR to ensure full coverage after a<br />

few additions <strong>and</strong> to avoid both precipitation <strong>of</strong> the nanohybrid<br />

<strong>and</strong> the generation <strong>of</strong> free gold NPs. An initial [Au seed]/[protein]<br />

MR=50wasfoundtobetheoptimalone<strong>and</strong>wasfurtherused<br />

during the sequential additions <strong>of</strong> Au ions to achieve the<br />

formation <strong>of</strong> the metallic nanowires (see Table 1).<br />

The morphologies <strong>of</strong> the resulting products were investigated<br />

by transmission electron microscopy (TEM) <strong>and</strong> scanning<br />

pubs.acs.org/JPCL<br />

Figure 1. TEM images <strong>of</strong> (a) bare Lys fibrils <strong>and</strong> Au-fibril biohybrids obtained after (b) zero, (c) one, (d) two, (e) three, <strong>and</strong> (f) four sequential<br />

additions <strong>of</strong> the Au growth solution; (g) bright field <strong>and</strong> (h) dark field TEM images <strong>of</strong> fully covered fibrils; (i) SEM image <strong>of</strong> a Au-metallic<br />

biohybrid nanowire. Scale bars are (a,i) 500, (b) 100, <strong>and</strong> (c-h) 200 nm.<br />

Table 1. Au Content <strong>of</strong> the Obtained Nanohybrids<br />

sample ([Au]/[protein] MR) Au concentration (10 -5 M)<br />

biohybrid 1(50) 3.4<br />

biohybrid 2 (100) 3.9<br />

biohybrid 3 (125) 4.6<br />

biohybrid 4 (150) 5.3<br />

biohybrid 5 (200) 6.0<br />

transmission microscopy (SEM). Figure 1a shows a typical<br />

TEM image <strong>of</strong> the as-prepared Lys fibrils, with lengths<br />

ranging from 0.4 to 8 μm <strong>and</strong> widths <strong>of</strong> 12-20 nm. The<br />

metallic coverage <strong>of</strong> the nan<strong>of</strong>ibrils could be easily controlled<br />

by changing the number <strong>of</strong> sequential additions <strong>of</strong><br />

gold growth solution into the fibril solution over already<br />

preformed Au nanocrystals. Figure 1b-e shows the progressive<br />

coverage <strong>of</strong> the nan<strong>of</strong>ibrils after increasing the number<br />

<strong>of</strong> sequential additions, with the formation <strong>of</strong> more metallic<br />

nanoparticles onto the fibril surface. TEM revealed that<br />

metallic nanoparticles on the fibril surfaces possess a size<br />

ranging from 3.7 to 8.0 ( 2.0 nm, depending on the number<br />

<strong>of</strong> sequential additions <strong>of</strong> the gold growth solution. The<br />

continuous coverage is a result <strong>of</strong> the growth <strong>of</strong> pre-existing<br />

NPs <strong>and</strong> generation <strong>of</strong> new ones on the fibril surface <strong>and</strong><br />

subsequent fusion <strong>of</strong> adjacent as-synthesized NPs. The fully<br />

covered fibrils display a relatively rougher surface than<br />

pristine <strong>and</strong> partially covered fibrils, as can be observed<br />

from the TEM image obtained in the backscattering mode<br />

<strong>and</strong> the SEM picture obtained without the need for metallic<br />

contrast (see Figure 1h,i). Also, the size <strong>of</strong> the metallic NPs<br />

clearly increases, <strong>and</strong> the size population can also broadened<br />

(see Figure S1, Supporting Information). In addition,<br />

r XXXX American Chemical Society 2681 DOI: 10.1021/jz101029f |J. Phys. Chem. Lett. 2010, 1, 2680–2687<br />

200


Figure 2. (a) XRD pattern <strong>of</strong> spherical Au NPs on the surface <strong>of</strong> protein fibrils; (b) EDX pattern <strong>of</strong> a gold hybrid fibril.<br />

under conditions <strong>of</strong> excessive addition <strong>of</strong> Au ions in solution,<br />

an increasing presence <strong>of</strong> fibril networks is observed as a<br />

consequence <strong>of</strong> fibril bundling, which leads to precipitation<br />

<strong>of</strong> the nanowires after several days (see Figure S2, Supporting<br />

Information).<br />

The formation <strong>of</strong> hybrid fibrils was further characterized by<br />

energy dispersive X-ray (EDX) spectroscopy <strong>and</strong> X-ray diffraction<br />

(XRD). XRD patterns <strong>of</strong> the resulting hybrid fibrils display<br />

the classical peaks corresponding to metallic spherical gold<br />

NPs. The crystallite sizes calculated according to the modified<br />

Scherrer relation 55 for the hybrid fibrils range from 5 to 8 nm<br />

depending on the number <strong>of</strong> additions <strong>of</strong> Au growth solution.<br />

The EDX spectrum (Figure 2b) <strong>of</strong> protein-fibril-supported Au<br />

NPs shows the peaks corresponding to C, O, S, <strong>and</strong> Au,<br />

confirming the existence <strong>of</strong> Au NPs on the surface <strong>of</strong> the Lys<br />

fibrils. The S peak is characteristic <strong>of</strong> protein samples. 56<br />

To study the performance <strong>of</strong> the Au NP hybrid fibrils as<br />

catalysts, we have chosen the borohydride reduction <strong>of</strong><br />

p-nitrophenol as a model reaction. It is well documented that<br />

this reaction can be catalyzed by noble metal nanoparticles,<br />

<strong>and</strong> the color changes involved in the reduction also provide a<br />

simple way based on spectroscopic measurements for monitoring<br />

the reaction kinetics. 57,58 Under a neutral or acidic<br />

condition, p-nitrophenol solution exhibits a strong absorption<br />

peak at 317 nm. Upon the addition <strong>of</strong> NaBH4, the alkalinity <strong>of</strong><br />

pubs.acs.org/JPCL<br />

the solution increases, <strong>and</strong> p-nitrophenolate ions would<br />

become the dominating species, together with a spectral shift<br />

to 400 nm for the absorption peak. 59,60 This peak remains<br />

unaltered with time, suggesting that the reduction did not take<br />

place in the absence <strong>of</strong> a catalyst, as reported elsewhere. 61,62<br />

However, the addition <strong>of</strong> a small amount <strong>of</strong> the Au-based<br />

catalyst (i.e., metal hybrid fibrils with a Au concentration<br />

ranging from 0.002 to 0.02 mM) to the above reaction<br />

mixture causes fading <strong>and</strong> ultimate bleaching <strong>of</strong> the yellow<br />

color <strong>of</strong> the reaction mixture quickly. Time-dependent absorption<br />

spectra <strong>of</strong> this reaction mixture show the disappearance<br />

<strong>of</strong> the peak at 400 nm that is accompanied by a gradual<br />

development <strong>of</strong> a new peak at 300 nm corresponding to the<br />

formation <strong>of</strong> p-aminophenol (Figure 3a). The UV spectra also<br />

exhibit an isosbestic point between two absorption b<strong>and</strong>s,<br />

indicating that only two principal species, p-nitrophenol <strong>and</strong><br />

p-aminophenol, influence the reaction kinetics. Similar spectral<br />

changes were also observed with other hybrid fibrils with<br />

different Au NPconcentrations. These results indicate that the<br />

hybrid fibrils can successfully catalyze the reduction reaction.<br />

Several research groups have also reported similar kinds <strong>of</strong><br />

spectral changes <strong>of</strong> p-nitrophenol during its catalytic reduction<br />

in the presence <strong>of</strong> both supported <strong>and</strong> unsupported<br />

spherical metallic nanoparticles. 63-65 In this reduction process,<br />

as the concentration <strong>of</strong> NaBH4 in the reaction mixture far<br />

r XXXX American Chemical Society 2682 DOI: 10.1021/jz101029f |J. Phys. Chem. Lett. 2010, 1, 2680–2687<br />

201


pubs.acs.org/JPCL<br />

Figure 3. (a) Time evolution <strong>of</strong> UV-vis spectra <strong>and</strong> (b) variation <strong>of</strong> absorbance with time upon conversion <strong>of</strong> p-nitrophenol to<br />

p-aminophenol in the presence <strong>of</strong> biohybrid 2 with Au concentrations <strong>of</strong> (0) 0.26, (b) 0.53, (4) 0.79, (1) 1.05, <strong>and</strong> (þ) 1.31 10 -5 M.<br />

The arrow in (a) denotes the decreasing presence <strong>of</strong> p-nitrophenol in the mixture as the reduction reaction proceeds. Variation <strong>of</strong> (c) k app<br />

<strong>and</strong> (d) k nor with Au concentration present in the reduction reaction: (9) biohybrid 1, (O) biohybrid 2, (2) biohybrid 3,; (þ) biohybrid 4, <strong>and</strong><br />

(3) biohybrid 5.<br />

exceeds the concentration <strong>of</strong> p-nitrophenol, the reaction rate<br />

is assumed to follow first-order kinetics. Keeping this in mind,<br />

we decided to measure the concentrations <strong>of</strong> p-nitrophenolate<br />

ions <strong>and</strong> thus to monitor the progress or kinetics <strong>of</strong> the<br />

reaction by recording the absorbance at 400 nm because the<br />

latter peak was much stronger than that at 315 nm. We<br />

calculated the values <strong>of</strong> the apparent rate constants (kapp) <strong>of</strong><br />

the catalytic reaction in the presence <strong>of</strong> different concentrations<br />

<strong>of</strong> the hybrid fibrils possessing different degrees <strong>of</strong> Au<br />

coverage from the change <strong>of</strong> absorbance with time (Figures 3b<br />

<strong>and</strong> S3, Supporting Information). These plots are straight<br />

lines, indicating that the reduction reaction effectively follows<br />

first-order kinetics. The k app values <strong>of</strong> the hybrid fibrils<br />

rise linearly with an increase <strong>of</strong> their concentration (<strong>and</strong>,<br />

thus, also with increases in Au concentration) in the reaction<br />

medium until reaching a maximum to decrease further (see<br />

Figure 3c). In addition, values <strong>of</strong> k app for p-nitrophenol<br />

reduction using hybrid fibrils were ∼1.5-4 times larger<br />

than those reported for glucose-reduced Au nanoparticle<br />

networks, 64 poly(amidoamine) (PAMAM)-dendrimer-supported<br />

spherical nanoparticles, 59 solid 1D Au nanobelts 66<br />

<strong>and</strong> nanorods, 67 polymer-micelle-supported Au NPs, 68,69 multicomponent<br />

microgels, 70 <strong>and</strong> 1D assemblies <strong>of</strong> Au NPs, 71 <strong>and</strong><br />

they were <strong>of</strong> similar order as those obtained for porous 1D<br />

nanobelts 66 or Au yolk-shell NPs. 72 Nevertheless, we have to<br />

note that we used very small amounts <strong>of</strong> gold catalysts in our<br />

experiments.<br />

Thus, to get a clearer picture <strong>of</strong> the efficacy <strong>of</strong> the composite<br />

material as a catalyst <strong>and</strong> also to compare the catalytic<br />

activity between hybrid fibrils with different degrees <strong>of</strong> coverage,<br />

the catalytic activities are also presented in terms <strong>of</strong> a<br />

normalized rate constant (knor) obtained by normalization<br />

<strong>of</strong> k app to the total amount (mmol) <strong>of</strong> catalyst (gold) used in<br />

the reaction. Figure 3d shows that the values <strong>of</strong> k nor for<br />

p-nitrophenol reduction in the presence <strong>of</strong> the different types<br />

<strong>of</strong> hybrid fibrils prepared at different [Au]/[protein] MR values<br />

are also different. In particular, k nor remains relatively constant<br />

from hybrid 1 to 3, that is, as the coverage <strong>of</strong> the fibrils<br />

increases, while for hybrids 4 <strong>and</strong> 5, k nor progressively<br />

reduces. This behavior might be a consequence <strong>of</strong> (i)<br />

changes in nanoparticle sizes on the fibril surfaces <strong>and</strong> (ii)<br />

a decrease in the available metal surface area for the<br />

catalytic reaction to take place as the fibril coverage is almost<br />

complete. In this regard, it is well-known that the catalytic<br />

activity increases as the size <strong>of</strong> Au NPs decreases because <strong>of</strong><br />

the increase <strong>of</strong> effective surface area. In our case, although<br />

the crystallite size increases as the concentration <strong>of</strong> the<br />

growth solution increases, we have noted a larger increase<br />

in nanoparticle mean size accompanied with an important<br />

polydispersity for nanohybrids 4 <strong>and</strong> 5, as previously<br />

r XXXX American Chemical Society 2683 DOI: 10.1021/jz101029f |J. Phys. Chem. Lett. 2010, 1, 2680–2687<br />

202


(15) Banholzer, M. J.; Millstone, J. E.; Qin, L.; Mirkin, C. A. Rationally<br />

Designed Nanostructures for Surface-Enhanced Raman<br />

Spectroscopy. Chem. Soc. Rev. 2008, 37, 885–897.<br />

(16) Li, W.; Camargo, P. H. C.; Lu, X.; Xia, Y. Dimers <strong>of</strong> Silver<br />

Nanospheres: Facile Synthesis <strong>and</strong> Their Use as Hot Spots for<br />

Surface-Enhanced Raman Scattering. Nano Lett. 2009, 9,<br />

485–490.<br />

(17) Rodríguez-Lorenzo, L.; Alvarez-Puebla, R. A.; Pastoriza-Santos,<br />

I.; Mazzucco, S.; Stephan, O.; Kociak, M.; Liz-Marzan, L. M.;<br />

García de Abajo, F. J. Zeptomol Detection Through Controlled<br />

Ultrasensitive Surface-Enhanced Raman Scattering. J. Am.<br />

Chem. Soc. 2009, 131, 4616–4618.<br />

(18) Alvarez-Puebla, R. A.; Liz-Marzan, L. M. SERS-Based Diagnosis<br />

<strong>and</strong> Biodetection. Small 2010, 6, 604–610.<br />

(19) van Bommel, K. J. C.; Friggeri, A.; Shinkai, S. Organic<br />

Templates for the Generation <strong>of</strong> Inorganic Materials. Angew.<br />

Chem., Int. Ed. 2003, 42, 980–999.<br />

(20) Baron, R.; Willner, B.; Willner, I. Biomolecule-Nanoparticle<br />

Hybrids as Functional Units for Nanobiotechnology. Chem.<br />

Commun. 2007, 323–332.<br />

(21) Ofir, Y.; Samanta, B.; Rotello, V. M. Polymer <strong>and</strong> Biopolymer<br />

Mediated <strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> Gold Nanoparticles. Chem. Soc.<br />

Rev. 2008, 37, 1814–1825.<br />

(22) Sotiropoulou, S.; Sierra-Sastre, Y.; Mark, S. S.; Batt, C. A.<br />

Biotemplated Nanostructured Materials. Chem. Mater. 2008,<br />

20, 821–834.<br />

(23) Behrens, S. S. Synthesis <strong>of</strong> Inorganic Nanomaterials<br />

Mediated by Protein Assemblies. J. Mater. Chem. 2008, 18,<br />

3788–3798.<br />

(24) Warner, M. G.; Hutchison, J. E. Linear Assemblies <strong>of</strong> Nanoparticles<br />

Electrostatically Organized on DNA Scaffolds. Nat.<br />

Mater. 2003, 2, 272–277.<br />

(25) Gun, Q.; Chen, C.; Haynie, D. T. Cobalt Metallization <strong>of</strong> DNA:<br />

Toward Magnetic Nanowires. Nanotechnology 2005, 16,<br />

1358–1363.<br />

(26) Kundu, S.; Liang, H. Photochemical Synthesis <strong>of</strong> Electrically<br />

Conductive CdS Nanowires on DNA Scaffolds. Adv. Mater.<br />

2008, 20, 826–831.<br />

(27) Wirges, C. T.; Timper, J.; Fischler, M.; Sologubenko, A. S.;<br />

Mayer, J.; Simon, U.; Carell, T. Controlled Nucleation <strong>of</strong> DNA<br />

Metallization. Angew. Chem., Int. Ed. 2009, 48, 219–223.<br />

(28) Knez, M.; Bittner, A. M.; Boes, F.; Wege, C.; Jeske, H.; Maiβ, E.;<br />

Kern, K. Biotemplate Synthesis <strong>of</strong> 3-nm Nickel <strong>and</strong> Cobalt<br />

Nanowires. Nano Lett. 2003, 3, 1079–1082.<br />

(29) Dujardin, E.; Peet, C.; Stubbs, G.; Culver, J.; Mann, S. Organization<br />

<strong>of</strong> Metallic Nanoparticles Using Tobacco Mosaic Virus<br />

Templates. Nano Lett. 2003, 3, 413–417.<br />

(30) Tsukamoto, R.; Muraoka, M.; Seki, M.; Tabata, H.; Yamashita,<br />

I. Synthesis <strong>of</strong> CoPt <strong>and</strong> FePt3 Nanowires Using the Central<br />

Channel <strong>of</strong> Tobacco Mosaic Virus as a Biotemplate. Chem.<br />

Mater. 2007, 19, 2389–2391.<br />

(31) Balci, S.; Leinberger, D. M.; Knez, M.; Bittner, A. M.; Boes, F.;<br />

Kadri, A.; Wege, C.; Jeske, H.; Kern, K. Printing <strong>and</strong> Aligning<br />

Mesoscale Patterns <strong>of</strong> Tobacco mosaic Virus on Surfaces. Adv.<br />

Mater. 2008, 20, 2195–2200.<br />

(32) Fu, X.; Wang, Y.; Huang, L.; Sha, Y.; Gui, L.; Lai, L.; Tang, Y.<br />

Assemblies <strong>of</strong> Metal Nanoparticles <strong>and</strong> <strong>Self</strong>-Assembled Peptide<br />

Fibrils ; Formation <strong>of</strong> Double Helical <strong>and</strong> Single-Chain<br />

Arrays <strong>of</strong> Metal Nanoparticles. Adv. Mater. 2003, 15,902–906.<br />

(33) Gazit, E. <strong>Self</strong>-Assembled Peptide Nanostructures: the Design<br />

<strong>of</strong> Molecular Building Blocks <strong>and</strong> Their Technological Utilization.<br />

Chem. Soc. Rev. 2007, 36, 1263–1269.<br />

(34) Sharma, N.; Top, A.; Kiick, K. L.; Pochan, D. J. One-Dimensional<br />

Gold Nanoparticle Arrays by Electrostatically Directed<br />

pubs.acs.org/JPCL<br />

Organization Using Polypeptide <strong>Self</strong>-<strong>Assembly</strong>. Angew.<br />

Chem., Int. Ed. 2009, 48, 7078–7082.<br />

(35) Scheibel, T.; Parthasarathy, R.; Sawicki, G.; Lin, X. M.; Jaeger,<br />

H.; Lindquist, S. L. Conducting Nanowires Built by Controlled<br />

<strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> Amyloid Fibers <strong>and</strong> Selective Metal Deposition.<br />

Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 4527–4532.<br />

(36) Behrens, S.; Wu, J.; Habitch, W.; Unger, E. Silver Nanoparticle<br />

<strong>and</strong> Nanowire Formation by Microtubule Templates. Chem.<br />

Mater. 2004, 16, 3085–3090.<br />

(37) Cherny, I.; Gazit, E. Amyloids: Not Only Pathological Agents<br />

but Also Ordered Nanomaterials. Angew. Chem., Int. Ed. 2008,<br />

47, 4062–4069.<br />

(38) Slocik, J. M.; Kim, S. N.; Whitehead, T. A.; Clark, D. S.; Naik,<br />

R. R. Biotemplated Metal Nanowires Using Hyperthermophilic<br />

Protein Filaments. Small 2009, 5, 2038–2042.<br />

(39) Bigall, N. C.; Reitzing, M.; Naumann, W.; Simon, P.; van Pee,<br />

K.-H.; Eychm€uller, A. Fungal Templates for Noble-Metal<br />

Nanoparticles <strong>and</strong> Their Application in Catalysis. Angew.<br />

Chem., Int. Ed. 2008, 47, 7876–7879.<br />

(40) Retches, M.; Gazit, E. Casting Metal Nanowires Within Discrete<br />

<strong>Self</strong>-Assembled Peptide Nanotubes. Science 2003, 300,<br />

625–627.<br />

(41) Kinsella, J. M.; Ivanisevic, A. Enzymatic Clipping <strong>of</strong> DNA<br />

Wires Coated with Magnetic Nanoparticles. J. Am. Chem.<br />

Soc. 2005, 127, 3276–3277.<br />

(42) Bai, H.; Xu, K.; Xu, Y.; Matsui, H. Fabrication <strong>of</strong> Au Nanowires<br />

<strong>of</strong> Uniform Length <strong>and</strong> Diameter Using a Monodisperse <strong>and</strong><br />

Rigid Biomolecular Template: Collagen-like Triple Helix. Angew.<br />

Chem., Int. Ed. 2007, 46, 3319–3322.<br />

(43) Ostrov, N.; Gazit, E. Genetic Engineering <strong>of</strong> Biomolecular<br />

Scaffolds for the Fabrication <strong>of</strong> Organic <strong>and</strong> Metallic Nanowires.<br />

Angew. Chem., Int. Ed. 2010, 49, 3018–3021.<br />

(44) Guha, S.; Banerjee, A. <strong>Self</strong>-Assembled Robust Dipeptide<br />

Nanotubes <strong>and</strong> Fabrication <strong>of</strong> Dipeptide-Capped Gold Nanoparticles<br />

on the Surface <strong>of</strong> these Nanotubes. Adv. Funct. Mater.<br />

2009, 19, 1949–1961.<br />

(45) Ryu, J.; Lim, S. Y.; Park, C. B. Photoluminescent Peptide<br />

Nanotubes. Adv. Mater. 2009, 11, 1577–1581.<br />

(46) Gorzny, M. L.; Walton, A. S.; Evans, S. D. Synthesis <strong>of</strong> High-<br />

Surface-Area Platinum Nanotubes Using a Viral Template.<br />

Adv. Funct. Mater. 2010, 20, 1295–1300.<br />

(47) Balci, S.; Noda, K.; Bittner, A.; Kadri, A.; Wege, C.; Jeske, H.;<br />

Kern, K. <strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> Metal-Virus Nanodumbbells. Angew.<br />

Chem., Int. Ed. 2007, 46, 3149–3151.<br />

(48) Dobson, C. M. Protein Folding <strong>and</strong> Misfolding. Nature 2003,<br />

426, 884–890.<br />

(49) Jahn, T. R.; Radford, S. E. Folding Versus Aggregation: Polypeptide<br />

Conformations on Competing Pathways. Arch. Biochem.<br />

Biophys. 2008, 469, 100–117.<br />

(50) Frare, E.; De Laureto, P. P.; Zurdo, J.; Dobson, C. M.; Fontana,<br />

A. A Highly Amyloidogenic Region <strong>of</strong> Hen Lysozyme. J. Mol.<br />

Biol. 2004, 340, 1153–1165.<br />

(51) Pawar, A. P.; Dubai, K. F.; Zurdo, J.; Chiti, F.; Vendruscolo,<br />

M.; Dobson, C. M. Prediction <strong>of</strong> “Aggregation-Prone” <strong>and</strong><br />

“Aggregation-Susceptible” Regions in Proteins Associated<br />

with Neurodegenerative Diseases. J. Mol. Biol. 2005, 350,<br />

379–392.<br />

(52) Lieu, V. H.; Wu, J. W.; Wang, S. S. S.; Wu, C. H. Inhibition<br />

<strong>of</strong> Amyloid Fibrillization <strong>of</strong> Hen Egg-White Lysozymes by<br />

Rifampicin <strong>and</strong> p-Benzoquinone. Biotechnol. Prog. 2007, 23,<br />

698–706.<br />

(53) Yu, L.; Banerjee, I. A.; Shima, A.; Rajan, K.; Matsui, H. Size-<br />

Controlled Ni Nanocrystal Growth on Peptide Nanotubes <strong>and</strong><br />

Their Magnetic Properties. Adv. Mater. 2004, 16, 709–712.<br />

r XXXX American Chemical Society 2686 DOI: 10.1021/jz101029f |J. Phys. Chem. Lett. 2010, 1, 2680–2687<br />

205


(54) Almeida, N. L.; Oliveira, C. L. P.; Torriani, I. L.; Loh, W.<br />

Calorimetric <strong>and</strong> Structural Investigation <strong>of</strong> the Interaction<br />

<strong>of</strong> Lysozyme <strong>and</strong> Bovine Serum Albumin with Poly(Ethylene<br />

Oxide) <strong>and</strong> its Copolymers. Colloids Surf., B 2004, 38,67–76.<br />

(55) Dutta, P.; Manivannan, A.; Seehra, M. S.; Shah, N.; Huffman,<br />

G. P. Magnetic Properties <strong>of</strong> Nearly Defect-Free Maghemite<br />

Nanocrystals. Phys. Rev. B 2004, 70, 174428/1–174428/7.<br />

(56) Shiomi, T.; Tsunoda, T.; Kawai, A.; Mizukami, F.; Sakaguchi, K.<br />

Biomimetic Synthesis <strong>of</strong> Lysozyme-Silica Hybrid Hollow<br />

Particles Using Sonochemical Treatment: Influence <strong>of</strong> pH<br />

<strong>and</strong> Lysozyme Concentration on Morphology. Chem. Mater.<br />

2007, 19, 4486–4493.<br />

(57) Schrinner, M.; Ballauff, M.; Talmon, Y.; Kauffmann, Y.; Thun,<br />

J.; Moller, M.; Breu, J. Dynamical Quorum Sensing <strong>and</strong><br />

Synchronization in Large Populations <strong>of</strong> Chemical Oscillators.<br />

Science 2009, 323, 617–620.<br />

(58) Qin, G.; W. Pei, W.; Ma, X.; Xu, X.; Ren, Y.; Sun, W.; Zuo, L.<br />

Enhanced Catalytic Activity <strong>of</strong> Pt Nanomaterials: From<br />

Monodisperse Nanoparticles to <strong>Self</strong>-Organized Nanoparticle-Linked<br />

Nanowires. J. Phys. Chem. C 2010, 114, 6909–<br />

6913.<br />

(59) Hayakawa, K.; Yoshimura, T.; Esumi, K. Preparation <strong>of</strong> Gold-<br />

Dendrimer Nanocomposites by Laser Irradiation <strong>and</strong> Their<br />

Catalytic Reduction <strong>of</strong> 4-Nitrophenol, Immobilization <strong>and</strong><br />

Recovery <strong>of</strong> Au Nanoparticles from Anion Exchange Resin:<br />

Resin-Bound Nanoparticle Matrix as a Catalyst for the Reduction<br />

<strong>of</strong> 4-Nitrophenol. Langmuir 2003, 19, 5517–5521.<br />

(60) Praharaj, S.; Nath, S.; Gosh, S. K.; Kundu, S.; Pal, T. Immobilization<br />

<strong>and</strong> Recovery <strong>of</strong> Au Nanoparticles from Anion Exchange<br />

Resin: Resin-Bound Nanoparticle Matrix as a Catalyst<br />

for the Reduction <strong>of</strong> 4-Nitrophenol. Langmuir 2004, 20,<br />

9889–9892.<br />

(61) Rashid, M. H.; Bhattacharjee, R. R.; Kotal, A.; M<strong>and</strong>al, T. K.<br />

Synthesis <strong>of</strong> Spongy Gold Nanocrystals with Pronounced<br />

Catalytic Activities. Langmuir 2006, 22, 7141–7143.<br />

(62) Rashid, M. H.; M<strong>and</strong>al, T. K. Organic Lig<strong>and</strong>-Mediated Synthesis<br />

<strong>of</strong> Shape-Tunable Gold Nanoparticles: An Application <strong>of</strong><br />

Their Thin Film as Refractive Index Sensors. J. Phys. Chem. C<br />

2007, 111, 9684–9693.<br />

(63) Esumi, K.; Isono, R.; Yoshimura, T. Preparation <strong>of</strong> PAMAM<strong>and</strong><br />

PPI-Metal (Silver, Platinum, <strong>and</strong> Palladium) Nanocomposites<br />

<strong>and</strong> Their Catalytic Activities for Reduction <strong>of</strong><br />

4-Nitrophenol. Langmuir 2004, 20, 237–243.<br />

(64) Liu, J.; Qin, G.; Raavendran, P.; Ikushima, Y. Facile “Green”<br />

Synthesis, Characterization, <strong>and</strong> Catalytic Function <strong>of</strong> β-D-<br />

Glucose-Stabilized Au Nanocrystals. Chem.;Eur. J. 2006, 12,<br />

2131–2138.<br />

(65) Chen, X.; Zhao, D.; An, Y.; Shi, L.; Hou, W.; Chen, L. Catalytic<br />

Properties <strong>of</strong> Gold Nanoparticles Immobilized on the Surfaces<br />

<strong>of</strong> Nanocarriers. J. Nanopart. Res. 2010, 12, 1877–1887.<br />

(66) Li, L.; Wang, Z.; Huang, T.; Xie, J.; Qin, L. Porous Gold<br />

Nanobelts Templated by Metal-Surfactant Complex Nanobelts.<br />

Langmuir 2010, 26, 12330–12335.<br />

(67) Bai, X.; Gao, Y.; Liu, H.-g.; Zheng, L. Synthesis <strong>of</strong> Amphiphilic<br />

Ionic Liquids Terminated Gold Nanorods <strong>and</strong> Their Superior<br />

Catalytic Activity for the Reduction <strong>of</strong> Nitro Compounds.<br />

J. Phys. Chem. C 2009, 113, 17730–17736.<br />

(68) Wang, Y.; Wei, G.; Zhang, W.; Jiang, X.; Zheng, P.; Shi, L.;<br />

Dong, A. Responsive Catalysis <strong>of</strong> Thermoresponsive Micelle-<br />

Supported Gold Nanoparticles. J. Mol. Catal. A: Chem. 2007,<br />

266, 233–238.<br />

(69) Chen, X.; An, Y.; Zhao, D.; He, Z.; Zhang, Y.; Cheng, J.; Shi, L.<br />

Core-Shell-Corona Au-Micelle Composites with a Tunable<br />

Smart Hybrid Shell. Langmuir 2008, 24, 8198–8204.<br />

pubs.acs.org/JPCL<br />

(70) Hain, J.; Schrinner, M.; Lu, Y.; Pich, A. Design <strong>of</strong> Multicomponent<br />

Microgels by Selective Deposition <strong>of</strong> Nanomaterials.<br />

Small 2008, 4, 2016–2024.<br />

(71) Huang, T.; Meng, F.; Qi, L. Facile Synthesis <strong>and</strong> One-Dimensional<br />

<strong>Assembly</strong> <strong>of</strong> Cyclodextrin-Capped Gold Nanoparticles<br />

<strong>and</strong> Their Applications in Catalysis <strong>and</strong> Surface-Enhanced<br />

Raman Scattering. J. Phys. Chem. C 2009, 113, 13636–13642.<br />

(72) Lee, J.; Park, L. C.; Song, H. A Nanoreactor Framework <strong>of</strong><br />

a Au@SiO 2 Yolk/Shell Structure for Catalytic Reduction <strong>of</strong><br />

p-Nitrophenol. Adv. Mater. 2008, 20, 1523–1528.<br />

(73) Rashid, M. H.; M<strong>and</strong>al, T. K. Templateless Synthesis <strong>of</strong><br />

Polygonal Gold Nanoparticles: An Unsupported <strong>and</strong> Reusable<br />

Catalyst with Superior Activity. Adv. Funct. Mater. 2008, 18,<br />

2261–2271.<br />

(74) Kumar, S. S.; Kumar, C. S.; Mathiyarasu, J.; Phani, K. L.<br />

Stabilized Gold Nanoparticles by Reduction Using 3,4-Ethylenedioxythiophene-polystyrenesulfonate<br />

in Aqueous Solutions:<br />

Nanocomposite Formation, Stability, <strong>and</strong> Application<br />

in Catalysis. Langmuir 2007, 23, 3401–3408.<br />

(75) Burda, C.; Chen, X.; Narayanan, R.; El-Sayed, M. A. Chemistry<br />

<strong>and</strong> Properties <strong>of</strong> Nanocrystals <strong>of</strong> Different Shapes. Chem.<br />

Rev. 2005, 105, 1025–1102.<br />

(76) Knubovets, T.; Osterhout, J. J.; Connolly, P. J.; Klibanov, A. M.<br />

Structure, Thermostability, And Conformational Flexibility <strong>of</strong><br />

Hen Egg-White Lysozyme Dissolved in Glycerol. Proc. Natl.<br />

Acad. Sci. U.S.A. 1999, 96, 1262–1267.<br />

r XXXX American Chemical Society 2687 DOI: 10.1021/jz101029f |J. Phys. Chem. Lett. 2010, 1, 2680–2687<br />

206


DOI: 10.1002/chem.201003679<br />

One-Dimensional Magnetic Nanowires Obtained by Protein Fibril<br />

Biotemplating<br />

JosuØ Juµrez, Adriana Cambón, Antonio Topete, Pablo Taboada,* <strong>and</strong><br />

Víctor Mosquera [a]<br />

Abstract: Magnetic nanowires were obtained<br />

through the in situ synthesis <strong>of</strong><br />

magnetic material by Fe-controlled<br />

nanoprecipitation in the presence <strong>of</strong><br />

two different protein (human serum albumin<br />

(HSA) <strong>and</strong> lysozyme (Lys)) fibrils<br />

as biotemplating agents. The<br />

structural characteristics <strong>of</strong> the biotemplates<br />

were transferred to the hybrid<br />

magnetic wires. They exhibited excel-<br />

Introduction<br />

A key issue in nanotechnology is the development <strong>of</strong> conceptually<br />

simple construction techniques for the mass fabrication<br />

<strong>of</strong> identical nanoscale structures. Interest in one-dimensional<br />

(1D) nanoscale materials <strong>and</strong> devices, <strong>of</strong>ten<br />

called nanowires, nanotubes or nanorods, has risen sharply<br />

in recent years. In particular, 1D magnetic nanostructures<br />

exhibit unique magnetic properties due to geometric confinement,<br />

magnetostatic interactions <strong>and</strong> nanoscale domain<br />

formation, [1] which endow them with potential applicability<br />

in data storage <strong>and</strong> logic devices, [2a] magneto-transport behaviour,<br />

[2b] micromechanical sensors [2c] <strong>and</strong> biomedicine. [2d]<br />

As a result <strong>of</strong> their high aspect ratios, 1D magnetic entities<br />

possess larger dipole moments than individual nanoparticles.<br />

This allows their manipulation with lower field strengths [3]<br />

<strong>and</strong> provides them with improved imaging contrast capabilities,<br />

which opens up the possibility <strong>of</strong> their use in new biomedical<br />

applications, specifically in the advent <strong>of</strong> low-field<br />

MRI. [4] Here we report the formation <strong>of</strong> 1D magnetic nanowires<br />

<strong>and</strong> assemblies <strong>of</strong> magnetic nanoparticles over linear<br />

nanosized biopolymer templates formed by spontaneous fibrillation,<br />

under suitable conditions, <strong>of</strong> two proteins—<br />

human serum albumin (HSA) <strong>and</strong> lysozyme (Lys)—by in<br />

situ co-precipitation <strong>of</strong> iron under suitable conditions. These<br />

[a] J. Juµrez, A. Cambón, A. Topete, P. Taboada, V. Mosquera<br />

Grupo de Física de Coloides y Polímeros<br />

Departamento de Física de la Materia Condensada<br />

Facultad de Física, Universidad de Santiago de Compostela<br />

15782 Santiago de Compostela (Spain)<br />

Fax: (+ 34) 881814112<br />

E-mail: pablo.taboada@usc.es<br />

Supporting information for this article is available on the WWW<br />

under http://dx.doi.org/10.1002/chem.201003679.<br />

7366<br />

lent magnetic properties as a consequence<br />

<strong>of</strong> the 1D assembly <strong>and</strong> fusion<br />

<strong>of</strong> magnetite nanoparticles as ascertained<br />

by SQUID magnetometry.<br />

Prompted by these findings, we also<br />

Keywords: biotemplating · imaging<br />

agents · magnetic nanowires · nanostructures<br />

· protein fibrils<br />

checked their potential applicability as<br />

MRI contrast agents. The magnetic<br />

wires exhibited large r 2* relaxivities<br />

<strong>and</strong> sufficient contrast resolution even<br />

in the presence <strong>of</strong> an extremely small<br />

amount <strong>of</strong> Fe in the magnetic hybrids,<br />

which would potentially enable their<br />

use as T 2 contrast imaging agents.<br />

magnetic 1D nanostructures possess both high saturation<br />

magnetisations <strong>and</strong> spin–spin relaxivities at low magnetic<br />

fields, which makes them suitable c<strong>and</strong>idates for use as<br />

imaging contrast agents in MRI.<br />

There are many literature reports on the fabrication <strong>of</strong><br />

1D magnetic nanostructures, which can be assembled by (bio)templating,<br />

spontaneous self-assembly by magnetic dipolar<br />

interactions <strong>of</strong> nanoparticles, chemical synthesis, lithographic<br />

methods, laser etching or microcontact printing. [5]<br />

Owing to its relative simplicity, template-directed synthesis<br />

is one <strong>of</strong> the most attractive methods for preparing magnetic<br />

1D nanomaterials. [5a,b] This approach <strong>of</strong>fers exciting alternatives<br />

to the costly nanolithography-based “top-down” technologies<br />

or multi-step chemical procedures. Various linear<br />

nanometer-scale materials, including molecular dispersed<br />

<strong>and</strong> assembled polymers, [6] carbon nanotubes [5a,7] or biological<br />

scaffolds, have been employed to prepare 1D magnetic<br />

structures <strong>and</strong> assemblies. [8] In particular, the capability <strong>of</strong><br />

biopolymers to generate 1D magnetic nanostructures is<br />

really exciting because nature provides a renewable <strong>and</strong><br />

highly diverse source <strong>of</strong> nanometer-scale ordered complexes<br />

that can be used to replicate inorganic materials as well-defined<br />

structures. [9] Dextran, for example, has been used in<br />

the creation <strong>of</strong> elongated assemblies <strong>of</strong> spherical nanoparticles<br />

suitable for use as MRI contrast agents <strong>and</strong> tumour targeting<br />

species. [2d, 10] DNA was also used as a stabiliser for the<br />

formation <strong>of</strong> ordered nanowires <strong>of</strong> magnetic nanoparticles.<br />

[2b, 11] These nanowire assemblies resulted in stable magnetic<br />

fluids with high relaxivities at low fields useful for<br />

MRI imaging. The protein cage <strong>of</strong> cowpea chlorotic mottle<br />

virus (CCMV) was explored with the same goal, to incorporate<br />

high payloads <strong>of</strong> Gd 3+ with elevated molecular relaxivities,<br />

[12] <strong>and</strong> the formation either <strong>of</strong> magnetic nanotubes or <strong>of</strong><br />

magnetic nanowires with the aid <strong>of</strong> magnetic bacteria (Mag-<br />

2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2011, 17, 7366 – 7373


netospirillum magneticum strain AMB-1), [13] <strong>and</strong> tobacco<br />

mosaic virus [14] were also investigated.<br />

The spontaneous fibrillation undergone by different proteins<br />

under suitable conditions results in the formation <strong>of</strong><br />

amyloid-like fibrils, [15] potentially applicable as alternative<br />

templating agents for the construction <strong>of</strong> 1D magnetic materials.<br />

In particular, the proteins HSA <strong>and</strong> Lys are two proteins<br />

known to fibrillate under different solution conditions<br />

to form fibrils with diameters ranging from 10 to 50 nm <strong>and</strong><br />

lengths up to several microns [16] (see Figure S1 in the Supporting<br />

Information). The dimensions <strong>of</strong> the fibrils can be<br />

tuned by varying the solution conditions <strong>and</strong> the incubation<br />

times. They are also thermally stable at high temperatures<br />

<strong>and</strong> different pH values <strong>and</strong> they contain multiple potential<br />

ion-binding sites within their amino acid sequences, which<br />

enables their use in metallisation reactions under relatively<br />

harsh conditions. We have thus exploited these features to<br />

generate 1D magnetic nanoparticle assemblies <strong>and</strong> nanowires<br />

by in situ nanoparticle formation on the fibril surface<br />

by use <strong>of</strong> the electrostatic interaction between Fe 2 + <strong>and</strong><br />

Fe 3 + ions <strong>and</strong> the negatively charged amino acid residues located<br />

on the protein surfaces (mainly glutamic <strong>and</strong> aspartic<br />

acids) [17] <strong>and</strong> analysis <strong>of</strong> their magnetic properties <strong>and</strong> potential<br />

applicability as MRI contrast agents. Although biotemplating<br />

<strong>of</strong> metallic nanowires by peptides <strong>and</strong> protein fibrils<br />

has been reported previously, [8a, 18] less attention has<br />

been paid to the production <strong>of</strong> 1D protein fibril-templated<br />

magnetic structures, <strong>and</strong> their magnetic properties have not<br />

been analysed in detail. Bolaamphiphile peptide fibrils, for<br />

example, have been used to template the growth <strong>and</strong> size <strong>of</strong><br />

nickel nanocrystals on their surfaces <strong>and</strong> to tune their magnetic<br />

properties, [19] self-assembled peptides have also been<br />

conjugated to gadolinium ions, [20] <strong>and</strong> iron bionanomineralisation<br />

has been found to occur in human serum transferrin<br />

fibrils. [21]<br />

Results <strong>and</strong> Discussion<br />

Nanowire synthesis <strong>and</strong> characterisation: For in situ magnetic<br />

nanoparticle formation <strong>and</strong> subsequent fibril coverage,<br />

nanoparticle-coated protein fibrils were obtained by sequential<br />

addition <strong>of</strong> Fe 2 + <strong>and</strong> Fe 3 + ions (1:2 molar ratio) <strong>and</strong> ammonium<br />

hydroxide (NH 4OH) solutions to aqueous solutions<br />

containing self-assembled fibrils, with subsequent incubation<br />

at 80 8C (see the Experimental Section for protein fibrillation<br />

<strong>and</strong> nanoparticle growth details). The controlled precipitation<br />

<strong>of</strong> iron ions leads to the formation <strong>of</strong> magnetic nanoparticles<br />

on the fibrils (see Figure 1). The effects <strong>of</strong> iron<br />

concentration, iron/protein molar ratio, number <strong>of</strong> sequential<br />

additions <strong>and</strong>/or the presence <strong>of</strong> different reagents/stabilisers<br />

were tested to provide progressive full coverage <strong>of</strong><br />

the biotemplate <strong>and</strong> to generate a magnetic wire. A protocol<br />

composed <strong>of</strong> sequential additions <strong>of</strong> Fe ions in several steps<br />

(seeded-growth mechanism) was therefore performed (see<br />

the Experimental Section). A suitable [Fe]/ACHTUNGTRENUNG[protein] molar<br />

ratio (MR) was necessary to ensure full coverage after a<br />

FULL PAPER<br />

Figure 1. TEM images showing the progressive coverage <strong>of</strong> lysozyme fibrils<br />

to afford magnetic nanowires after a single addition <strong>of</strong> Fe ions at<br />

[Fe]/ACHTUNGTRENUNG[protein] molar ratios <strong>of</strong>: a) 25, b) 50, c) 100 <strong>and</strong> d) 100 together<br />

with one additional sequential addition <strong>of</strong> Fe ions at the same molar<br />

ratio. e) Size distributions <strong>of</strong> as-synthesised magnetic nanoparticles on<br />

the biotemplate surfaces. The dotted line represents the fit <strong>of</strong> the distribution.<br />

small number <strong>of</strong> addition steps <strong>and</strong> to avoid both precipitation<br />

<strong>of</strong> the nanohybrid <strong>and</strong> generation <strong>of</strong> free magnetic NPs.<br />

We should note that if the [Fe]/ACHTUNGTRENUNG[protein] molar ratio (MR =<br />

10, 25, 50) is low, complete fibril coverage is not achieved<br />

even after several sequential additions; in contrast, if the<br />

[Fe]/ACHTUNGTRENUNG[protein] MR is too large, bundling <strong>and</strong> precipitation <strong>of</strong><br />

the nanohybrids occurs (see Figure S2 in the Supporting Information).<br />

For these reasons an initial [Fe]/ACHTUNGTRENUNG[protein] MR=<br />

100 was found to be optimal <strong>and</strong> was further used for three<br />

sequential additions <strong>of</strong> Fe ions to achieve the formation <strong>of</strong><br />

the magnetic wires. Figure 1 a–d shows the progressive coverage<br />

<strong>of</strong> the nan<strong>of</strong>ibrils at different [Fe]/ACHTUNGTRENUNG[protein] molar<br />

ratios, with the formation <strong>of</strong> magnetic nanoparticles on the<br />

fibril surface. TEM revealed the primary magnetic nanoparticles<br />

to be fairly uniformly distributed on the biotemplate<br />

<strong>and</strong> to have sizes <strong>of</strong> about 6.8 2.0 nm with a relatively<br />

narrow size distribution (Figure 1 e), in fair agreement with<br />

those magnetic particles obtained in the absence <strong>of</strong> the protein<br />

template (see Figure S3 in the Supporting Information).<br />

Chem. Eur. J. 2011, 17, 7366 – 7373 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 7367


Also, it is necessary to note that fibril integrity <strong>and</strong> stability<br />

was perfectly maintained under these reaction conditions<br />

(see Figure S4 in the Supporting Information).<br />

With more addition steps carried out the fibrils became<br />

progressively covered with more magnetic material without<br />

any apparent increase in nanoparticle size. The continuous<br />

fibril coverage is a result <strong>of</strong> the fusion <strong>of</strong> adjacent as-synthesised<br />

magnetic nanoparticles (see Figure S5 in the Supporting<br />

Information). When a progressive <strong>and</strong> sufficient<br />

excess <strong>of</strong> Fe was added, the nanowire widths increased <strong>and</strong><br />

their surfaces also appeared rougher, which might indicate<br />

that more than one layer <strong>of</strong> magnetic material had been deposited<br />

onto the template (Figure 2 a–b). In this regard, the<br />

Figure 2. TEM images showing the progressive increases in the widths <strong>of</strong><br />

the magnetic wires <strong>and</strong> subsequent bundling after a) three, b) four or<br />

c) five sequential additions <strong>of</strong> Fe ions at MR = 100. d) After 5 sequential<br />

additions at MR = 100 in the presence <strong>of</strong> sodium citrate as stabiliser.<br />

widths <strong>of</strong> the magnetic wires could be modified from 6 to<br />

50 nm by increasing the number <strong>of</strong> sequential additions <strong>of</strong><br />

Fe ions onto the biotemplate. These values depend on the<br />

type <strong>of</strong> fibril used as the template, the [Fe]/ACHTUNGTRENUNG[protein] molar<br />

ratio <strong>and</strong> the number <strong>of</strong> sequential additions performed (see<br />

Figure 1 <strong>and</strong> Figure S2 in the Supporting Information). Nevertheless,<br />

constant width along the whole nan<strong>of</strong>ibrils could<br />

not be achieved, due to the non-uniform deposition <strong>of</strong> the<br />

ions onto the template. We are currently modifying the synthetic<br />

protocol with the goals both <strong>of</strong> improving uniformity<br />

in fibril coverage <strong>and</strong> <strong>of</strong> achieving perfect control over the<br />

magnetic nanowire dimensions <strong>and</strong> their resulting magnetic<br />

properties. In addition, under conditions <strong>of</strong> excessive Fe<br />

ions in solution an increasing presence <strong>of</strong> fibril networks as<br />

a consequence <strong>of</strong> fibril bundling was observed, leading to<br />

precipitation <strong>of</strong> the nanowires after several days (see Figure<br />

2 c). This can be a result either <strong>of</strong> magnetic interaction<br />

between nanowires or <strong>of</strong> inefficient stabilisation <strong>of</strong> the 1D<br />

7368<br />

P. Taboada et al.<br />

nanostructures. To prevent these phenomena, we decided to<br />

stabilise the magnetic nanowires further by the addition <strong>of</strong><br />

sodium citrate, which has been shown to be an effective stabiliser<br />

for water-soluble magnetic nanoparticles. [22] In this<br />

process, no fibril precipitation was observed for at least one<br />

month <strong>and</strong> a lower level <strong>of</strong> bundling was confirmed, probably<br />

as a consequence <strong>of</strong> the generation <strong>of</strong> a surface organic<br />

layer that moderately screens the magnetic interactions between<br />

magnetic covered fibrils (Figure 2 d).<br />

On the other h<strong>and</strong>, the morphologies <strong>and</strong> sizes <strong>of</strong> the<br />

fully covered fibrils made <strong>of</strong> HSA <strong>and</strong> Lys present some differences<br />

irrespective <strong>of</strong> the method used to create the magnetic<br />

nanowires. This is a consequence <strong>of</strong> the inherent structural<br />

differences between the fibrils assembled from the two<br />

proteins. In this respect, HSA-derived nanowires have shorter<br />

lengths (between 0.1 <strong>and</strong> 2 mm) <strong>and</strong> widths (4–10 nm),<br />

<strong>and</strong> have a curly morphology as a direct consequence <strong>of</strong> the<br />

template structure (see Figure S6 in the Supporting Information).<br />

[16a,b] In contrast, Lys-derived wires appear to be<br />

straighter <strong>and</strong> longer (lengths <strong>and</strong> widths <strong>of</strong> 0.4–8 mm <strong>and</strong><br />

12–20 nm, respectively). [16c,d] This observation additionally<br />

confirms that the overall structures <strong>of</strong> the fibril templates<br />

remain intact after surface modification.<br />

XRD analysis, selected area electron diffraction (SAED)<br />

patterns, high-resolution TEM (HR-TEM) images <strong>and</strong><br />

FTIR spectroscopic examination <strong>of</strong> the fully covered fibrils<br />

confirmed the formation <strong>and</strong> nature <strong>of</strong> the magnetic coverage<br />

on the template surface. The XRD pattern showed reflections<br />

indexed to (111), (220), (311), (222), (400), (422),<br />

(511), (440) <strong>and</strong> (533) planes, corresponding to the cubic<br />

spinel crystal structure <strong>of</strong> iron oxides (see Figure 3 a, JPCDS<br />

file No. 19-0629, magnetite, or JCPDS file 39-1346, maghemite).<br />

In addition, XRD demonstrated a polycrystalline<br />

structure with cell constant a 0 = 0.8381 nm, which is in good<br />

agreement with the st<strong>and</strong>ard card <strong>of</strong> iron oxides. The peak<br />

broadening <strong>of</strong> the XRD pattern also indicates that the resulting<br />

iron oxide crystallites were rather small. The crystallite<br />

sizes calculated from the modified Scherrer relation [23]<br />

for the magnetic fibrils <strong>of</strong> both Lys <strong>and</strong> HSA were about<br />

(6.2 2) nm, consistent with the size distributions measured<br />

from TEM images. SAED, HR-TEM <strong>and</strong> FTIR confirm the<br />

predominance <strong>of</strong> the magnetite phase on the biotemplate.<br />

The SAED pattern consisted <strong>of</strong> spots on the rings corresponding<br />

to the diffraction planes (111), (220), (311), (400),<br />

(511) <strong>and</strong> (440) characteristic <strong>of</strong> the magnetite phase, which<br />

also revealed the polycrystalline nature <strong>of</strong> the nanocomposites<br />

(see Figure 3 b). The difference between the polycrystalline<br />

electron diffraction patterns <strong>of</strong> magnetite <strong>and</strong> maghemite<br />

phases is solved by the presence <strong>of</strong> the (111) diffraction<br />

plane, distinctive <strong>of</strong> the magnetite fcc structure. [24] From the<br />

HR-TEM images (Figure 3 c) we can observe that the obtained<br />

magnetite nanoparticles are single crystalline, as indicated<br />

by the well-resolved lattice fringes. The distance between<br />

two adjacent planes is about 0.255 nm, which corresponds<br />

to the (311) lattice plane in the spinel structure <strong>of</strong><br />

Fe 3O 4. [25] The strong peak at 580 cm 1 in the FTIR plot (Figure<br />

3 d) additionally confirms that the phase is magnetite<br />

www.chemeurj.org 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2011, 17, 7366 – 7373


One-Dimensional Magnetic Nanowires<br />

Figure 3. a) XRD pattern <strong>and</strong> b) SAED pattern <strong>of</strong> the fully covered magnetic<br />

nanowires. c) HR-TEM image <strong>of</strong> a magnetite nanoparticle showing<br />

the interplanar distance; the FFT indicates that the nanoparticle is oriented<br />

in the [001] direction. d) FTIR spectrum <strong>of</strong> fully covered magnetic<br />

nanowires.<br />

rather than maghemite. The peaks at about 591 <strong>and</strong><br />

456 cm 1 in the spectrum indicate the presence <strong>of</strong> both magnetite<br />

<strong>and</strong> maghemite. [26]<br />

Magnetic <strong>and</strong> imaging contrast properties: To evaluate the<br />

potential use <strong>of</strong> these 1D hybrid magnetic nanostructures<br />

for bioimaging, we first investigated the magnetic properties<br />

<strong>of</strong> fully magnetite-covered hybrid lysozyme nan<strong>of</strong>ibrils with<br />

the aid <strong>of</strong> a superconducting quantum interference device<br />

(SQUID), <strong>and</strong> compared them with those <strong>of</strong> dispersed magnetic<br />

nanoparticles synthesised under the same solution conditions<br />

in the absence <strong>of</strong> protein fibrils. Figure 4 shows the<br />

field-dependent magnetisation <strong>of</strong> the fully covered fibrils<br />

<strong>and</strong> dispersed magnetic nanoparticles measured at 300 <strong>and</strong><br />

5 K, respectively. Their M/H curves at 300 K show nonlin-<br />

FULL PAPER<br />

Figure 4. Magnetisation versus applied magnetic field for: a) dispersed<br />

magnetite nanoparticles, <strong>and</strong> b) fully covered magnetic nanowires at 5 K<br />

(closed symbols) <strong>and</strong> 300 K (open symbols). c) Image showing the attraction<br />

<strong>of</strong> the nanowires to a magnet.<br />

ear, reversible characteristics with no hysteresis (zero coercivity),<br />

features <strong>of</strong> superparamagnetic behaviour, with saturation<br />

magnetisation (M s) values <strong>of</strong> about 60 <strong>and</strong> 47 emu g 1<br />

for fully covered <strong>and</strong> isolated nanoparticles, respectively.<br />

The observed magnetic behaviour is consistent with the<br />

common underst<strong>and</strong>ing that one-dimensionally stacked<br />

magnetic nanoparticles forming a uniform layer behave like<br />

single elongated particles, <strong>and</strong> so their cooperative response<br />

is more sensitive to a magnetic field. [27] Nevertheless, all the<br />

Chem. Eur. J. 2011, 17, 7366 – 7373 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 7369


above values are lower than the theoretical value <strong>of</strong> the<br />

bulk magnetite (92 emu g 1 ). [28] A decrease in M is <strong>of</strong>ten observed<br />

<strong>and</strong> is attributed to the surface contribution: spin<br />

canting, surface disorder, stoichiometry deviation, ion distribution<br />

<strong>and</strong> adsorbed species. [29] The elongated structures <strong>of</strong><br />

the nanowires apparently enhance the orientation <strong>of</strong> the<br />

magnetic moments <strong>of</strong> the individual nanoparticle constituents,<br />

increasing the net magnetisation. [13, 30] On the other<br />

h<strong>and</strong>, the magnetisation curves at 5 K show that all samples<br />

have hysteresis with coercitive fields (H c) <strong>of</strong> 352 <strong>and</strong><br />

270 Oe, remnant magnetisation <strong>of</strong> 16.5 <strong>and</strong> 10.2 emu g 1 <strong>and</strong><br />

M s values <strong>of</strong> about 90 <strong>and</strong> 56 emu g 1 for fully covered <strong>and</strong><br />

isolated nanoparticles, respectively. The magnetic properties<br />

<strong>of</strong> the hybrid fibrils were also tested by placing a magnet<br />

near the cuvette. The hybrid nanowires were completely attracted<br />

to the side <strong>of</strong> the cuvette nearest to the magnet as<br />

displayed in Figure 4 c, which additionally confirms the magnetic<br />

properties <strong>of</strong> the hybrid samples.<br />

Measurements <strong>of</strong> magnetisation versus temperature were<br />

performed by st<strong>and</strong>ard experimental zero-field-cooling<br />

(ZFC) <strong>and</strong> field-cooling (FC) protocols with the external<br />

magnetic field set at 500 Oe when necessary. Field-cooled<br />

(FC) magnetisation measurements provide information<br />

about coupling between the magnetic materials because a<br />

FC experiment suppresses thermal fluctuations <strong>and</strong> increases<br />

net magnetisation. The ZFC <strong>and</strong> FC magnetisation<br />

curves split below 150 <strong>and</strong> 195 K, <strong>and</strong> the ZFC magnetisation<br />

curves exhibit peaks around T = 35 <strong>and</strong> 105 K for fully<br />

covered fibrils <strong>and</strong> dispersed Fe 3O 4 nanoparticles, respectively<br />

(see Figure 5), identified as the blocking temperatures<br />

(T B). These data clearly show that there is a temperature<br />

range in which magnetic hybrid fibrils show collective ferromagnetic<br />

behaviour that is unique due to their aggregation<br />

status, as was also the case for nanoparticle chains. [5c,21a]<br />

Nevertheless, superparamagnetism remains after T B as a<br />

consequence <strong>of</strong> the anisotropic barrier blockage <strong>of</strong> the magnetisation<br />

orientation <strong>of</strong> the magnetic layer forming the<br />

hybrid fibril cooled with the ZFC process. [31] The larger T B<br />

<strong>and</strong> the less steep slope for magnetite nanoparticles in the<br />

above figure could be consistent with: i) a larger dipolar<br />

coupling, [27,32] or ii) larger size <strong>and</strong> anisotropy in the dispersions<br />

as a result <strong>of</strong> some degree <strong>of</strong> aggregation between the<br />

particles, which involves a larger number <strong>of</strong> nearest neighbouring<br />

particles than in the case <strong>of</strong> magnetic wires <strong>and</strong> a<br />

non-regular distribution <strong>of</strong> the particles in the aggregates, in<br />

contrast to what would be expected in the 1D nanowires.<br />

This implies that dipolar interactions will not equally affect<br />

all particles in the nanoparticle aggregates, giving rise to the<br />

observed broadening in T B.<br />

From the relationship T B=K aV/25k B (in which K a is the<br />

anisotropic constant, <strong>and</strong> k B the Boltzmann constant), for<br />

uniaxial non-interacting magnetite nanoparticles with a diameter<br />

<strong>of</strong> 10 nm, T B is found to be 20 K, in contrast to the<br />

values <strong>of</strong> about 35 K found for the nanowires <strong>and</strong> 105 K for<br />

dispersed nanoparticles, respectively. From this equation we<br />

can also derive an estimated value for the crystallite size in<br />

the hybrid fibrils <strong>and</strong> for the dispersed nanoparticles by use<br />

7370<br />

P. Taboada et al.<br />

Figure 5. ZFC-FC plots for: a) dispersed magnetite nanoparticles, <strong>and</strong><br />

b) fully covered magnetic nanowires at 5 K (closed symbols) <strong>and</strong> 300 K<br />

(open symbols). FC curves were obtained in the presence <strong>of</strong> a magnetic<br />

field <strong>of</strong> 500 Oe.<br />

<strong>of</strong> a K a value <strong>of</strong> 1.35 ”10 4 Jm 3 corresponding to bulk magnetisation.<br />

In this way, the estimated sizes are about 12 <strong>and</strong><br />

16 nm, larger than those calculated from XRD data. Possible<br />

reasons for these differences might be the wide size distribution<br />

populations, the magnetic dipole interactions between<br />

nanowires <strong>and</strong> nanoparticles, the magnetic texture<br />

formed in nanowires due to strong shape anisotropy, the aggregation<br />

state in the dispersed nanoparticles or the disordered<br />

surface spins. These factors can have strong influences<br />

on magnetic properties <strong>and</strong>, therefore, on calculated sizes.<br />

Additionally, the effective magnetic anisotropy constants <strong>of</strong><br />

Fe 3O 4 nanoparticles are normally less than its bulk value<br />

<strong>and</strong> vary with particle size. [29] All these factors lead to differences<br />

in the estimated particles sizes according to XRD <strong>and</strong><br />

magnetic data.<br />

To check the possible efficiency <strong>of</strong> the biohybrids as novel<br />

T 2 contrast agents, the relaxation times T 1 <strong>and</strong> T 2* were<br />

measured by NMR dispersion (NMRD) at 9.4 T (400 MHz<br />

proton Larmor frequency, 37 8C). The T 1 (spin–lattice) relaxation<br />

process is the result <strong>of</strong> the interaction between the excited<br />

nuclei <strong>and</strong> their surrounding environment, <strong>and</strong> the T 2*<br />

www.chemeurj.org 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2011, 17, 7366 – 7373


One-Dimensional Magnetic Nanowires<br />

(spin–spin) relaxation process <strong>of</strong> the interaction between the<br />

excited nuclei <strong>and</strong> those with lower energy. The efficiency <strong>of</strong><br />

an MRI contrast agent is commonly assessed in terms <strong>of</strong> its<br />

relaxivities (r 1 <strong>and</strong> r 2), which are rates <strong>of</strong> proton relaxation<br />

determined by Equation (1):<br />

1=T i,obs ¼ 1=T i,d þ r i½MŠ ðI ¼ 1, 2Þ ð1Þ<br />

in which 1/T i,obs is the observed solvent relaxation rate in the<br />

presence <strong>of</strong> a contrast agent, 1/T i,d is the relaxation rate <strong>of</strong><br />

the pure diamagnetic solvent, <strong>and</strong> [M] is the concentration<br />

<strong>of</strong> the contrast agent. [33] For a T 2 contrast agent, the higher<br />

the r 2*/r 1 ratio the better the contrast efficacy. The spin–spin<br />

<strong>and</strong> spin–lattice relaxation times weighted spin-echo MRI <strong>of</strong><br />

the fully-covered hybrid nan<strong>of</strong>ibrils are shown in Figure 6 a<br />

<strong>and</strong> Figure S7 in the Supporting Information, respectively.<br />

As shown in the former, the fully covered fibrils exhibited<br />

much stronger enhancement in T 2*-weighted relaxation<br />

times than the dispersed magnetic nanoparticles at the same<br />

Figure 6. a) T 2 relaxation rates as a function <strong>of</strong> iron concentration (mm<br />

Fe) for fully covered magnetic nanowires obtained by fibril biotemplating;<br />

r 2*=161 mm 1 s 1 .b)T 2*-weighted MR images <strong>of</strong> phantoms in the<br />

presence <strong>of</strong> different amounts <strong>of</strong> magnetic nanowires (expressed in Fe<br />

concentration). 1) 0.1”10 7 , 2) 0.2”10 7 , 3) 0.3”10 7 , 4) 0.5”10 7 ,5)1”<br />

10 7 <strong>and</strong> 6) 2”10 7 m.<br />

Fe concentration, with an r 2* value <strong>of</strong> 161 mm 1 s 1 (r 1=<br />

0.42 mm 1 s 1 ) <strong>and</strong> an r 2*/r 1 ratio <strong>of</strong> about 390. The effect <strong>of</strong><br />

the protein was found to be negligible. The former behaviour<br />

might be a result <strong>of</strong> the enhanced MR sensitivity due to<br />

clustering <strong>of</strong> individual magnetic nanocrystals on the fibrils.<br />

[34] Nevertheless, it is necessary to bear in mind that the<br />

large r 2*/r 1 ratio is a result <strong>of</strong> the different behaviour displayed<br />

by T 1 <strong>and</strong> T 2 with the applied magnetic field<br />

strength: both T 1 <strong>and</strong> T 2 generally increase at relatively low<br />

applied magnetic field, but T 1 reaches a maximum around<br />

0.5–1.0 T <strong>and</strong> then drastically decreases at larger magnetic<br />

fields. In contrast, T 2 remains almost constant or even slightly<br />

increases at larger applied magnetic fields, so the r 2*/r 1<br />

ratio definitely increases with applied magnetic field. [35]<br />

Thus, when comparing r 2*/r 1 ratios it is necessary to bear in<br />

mind the applied magnetic field at which the experiment<br />

was performed. Interestingly, the r 2* relaxivities <strong>of</strong> the<br />

hybrid fibrils are slightly larger than those obtained for<br />

DNA-templated magnetic nanowires, [11a] (bio)polymer-templated<br />

nanowires, [20,36] magnetic nanoworms [2d, 10b] or conventional<br />

superparamagnetic iron oxide nanoparticles. [37] Such a<br />

strong T 2 shortening effect, typically from stronger magnetic<br />

susceptibility, leads to spin dephasing <strong>and</strong> substantial MRI<br />

drop, which generated a “darkening” contrast, as seen in the<br />

T 2*-weighted MR images <strong>of</strong> the hybrids in phantoms measured<br />

at 9.4 T in a preclinical MRI. The mean T 2* values <strong>of</strong><br />

the phantoms get lower as the amount <strong>of</strong> fully covered<br />

nan<strong>of</strong>ibrils increases in solution (seen as black spots in the<br />

images) even though the amount <strong>of</strong> Fe present is extremely<br />

small, which avoids full darkness <strong>of</strong> the image (Figure 6 b).<br />

The increased MRI contrast observed for nanowires is the<br />

result <strong>of</strong> the increase in the susceptibility effects caused by<br />

the presence <strong>of</strong> Fe that can enhance spin–spin relaxation <strong>of</strong><br />

water molecules caused by the slightly larger magnetisation<br />

value [38] <strong>and</strong> the 1D assembly. [39]<br />

Conclusion<br />

FULL PAPER<br />

We report a method for obtaining magnetic Fe 3O 4 nanowires<br />

based on the use <strong>of</strong> HSA <strong>and</strong> Lys protein fibrils as bioscaffolds<br />

for generation <strong>of</strong> a complete magnetic Fe 3O 4<br />

coating layer on the biotemplate surface by controlled Fe<br />

nanoprecipitation in situ in the presence <strong>of</strong> NH 4OH. To achieve<br />

full coverage <strong>of</strong> the protein fibrils a suitable initial [Fe]/ACHTUNG-<br />

TRENUNG[protein] molar ratio must be selected <strong>and</strong> a suitable<br />

number <strong>of</strong> sequential additions <strong>of</strong> Fe ions must be performed.<br />

If these requirements are not fulfilled, partial coverage<br />

<strong>and</strong>/or protein bundling <strong>and</strong> further fibril precipitation<br />

are observed. Thanks to their large saturation magnetisation,<br />

the 1-dimensional magnetic nanowires can serve as efficient<br />

MRI contrast agents, with transverse relaxivities<br />

larger than those obtained previously for other 1D nanostructures<br />

derived by different methods. Their capability as<br />

contrast agents was verified by obtaining several MRI<br />

images at different Fe concentrations in phantoms. In spite<br />

<strong>of</strong> the very low concentrations used, the hybrid nanowires<br />

Chem. Eur. J. 2011, 17, 7366 – 7373 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 7371


exhibit an important contrast imaging effect, which confirms<br />

their potential use as MRI contrast agents. Future work will<br />

involve the refinement <strong>of</strong> the synthetic process to achieve<br />

perfect control <strong>of</strong> magnetic coverage, <strong>and</strong> thus hybrid fibril<br />

widths, magnetic nanowire lengths, population distribution<br />

<strong>and</strong> in vivo assays.<br />

Experimental Section<br />

Materials: Human serum albumin, lysozyme, FeCl2 <strong>and</strong> FeCl3 were obtained<br />

from Sigma Chemical. Proteins were used after further purification<br />

by liquid chromatography over a Superdex 75 column equilibrated<br />

with phosphate (0.01 m). All other chemicals were <strong>of</strong> the highest purity<br />

available. Experiments were carried out with doubly distilled, deionised<br />

<strong>and</strong> degassed water.<br />

Preparation <strong>of</strong> HSA <strong>and</strong> lysozyme fibril solutions: Protein fibril solutions<br />

were prepared by well-established protocols. [16] Briefly, HSA fibril stock<br />

solution was made by dissolving the protein to a final concentration typically<br />

<strong>of</strong> 2 mg mL 1 in phosphate buffer (pH 7.4, ionic strength 10 mm)<br />

<strong>and</strong> brought to pH 2.0 by addition <strong>of</strong> HCl. Lysozyme fibrils were obtained<br />

by forming a stock solution (2 mg mL 1 ) in glycine buffer (pH 2.0,<br />

ionic strength 10 mm). Both protein stock solutions were dialysed extensively<br />

against pure buffer. Protein concentrations were determined spectrophotometrically,<br />

with use <strong>of</strong> molar absorption coefficients <strong>of</strong> 35219<br />

<strong>and</strong> 37 609 m 1 cm 1 at 280 nm for HSA <strong>and</strong> Lys, respectively. [40] Before incubation,<br />

the solutions were filtered through a 0.2 mm filter into sterile<br />

test tubes. Samples were incubated at 65 8C in a reactor under reflux for<br />

10 <strong>and</strong> 3 days for HSA <strong>and</strong> Lys, respectively.<br />

Characterisation <strong>of</strong> the fibrils: Protein suspensions were applied to<br />

carbon-coated copper grids, blotted, washed, negatively stained with<br />

phosphotungstic acid (2%, w/v), air dried <strong>and</strong> then examined with a Phillips<br />

CM-12 transmission electron microscope operating at an accelerating<br />

voltage <strong>of</strong> 120 kV. Samples were diluted 20–200-fold when needed prior<br />

to deposition on the grids.<br />

Magnetic nanowire synthesis: A stock solution (typically 2 g L 1 ) <strong>of</strong><br />

either lysozyme or HSA fibrils was diluted 600-fold to avoid the presence<br />

<strong>of</strong> undesired aggregates during the synthesis <strong>of</strong> the magnetite nanowire.<br />

Different [Fe]/ACHTUNGTRENUNG[protein] molar ratios (10:1, 25:1 50:1, 100:1, 200:1 <strong>and</strong><br />

400:1) <strong>and</strong> sequential additions (one to five) <strong>of</strong> the iron salts were tested<br />

to enable the formation <strong>of</strong> magnetite nanoparticles on the biotemplate<br />

surface <strong>and</strong> to ensure full coverage <strong>of</strong> the fibril surface. In a typical synthesis,<br />

an aliquot (100 mL) <strong>of</strong> lysozyme fibril solution (2 gL 1 , 1.34 ”<br />

10 4 m) was placed in a flask reactor <strong>and</strong> diluted with doubly distilled, deionised<br />

water (60 mL). Fibril solution was left to incubate at 65 8C for 1 h<br />

with magnetic stirring. Separately, two aqueous solutions containing<br />

FeCl2 (25.4 mg, 0.1 m) <strong>and</strong> FeCl3, respectively, (64.9 mg, 0.1 m) were prepared<br />

<strong>and</strong> mixed in a 1:2 molar ratio. After the incubation period at<br />

658C, the temperature <strong>of</strong> the fibril solution was raised to 80 8C, <strong>and</strong> a<br />

specific volume <strong>of</strong> the iron salt mixed solution suitable to get the desired<br />

[Fe]/ACHTUNGTRENUNG[protein] molar ratio was then added under nitrogen. The solution<br />

was left to incubate for 3 h before the addition <strong>of</strong> NH4OH (1.5 m, 2 mL);<br />

the resulting solution was allowed to react for an additional 1 h <strong>and</strong> allowed<br />

to cool to room temperature. Stabilisation <strong>of</strong> the nanowires was<br />

additionally achieved by addition <strong>of</strong> sodium citrate solution (10 mm,<br />

1 mL) to the resulting nanowires at 80 8C <strong>and</strong> allowing to incubate for<br />

90 min. The resulting biohybrid solution was dialysed through a dialysis<br />

membrane (SpectraPor, Netherl<strong>and</strong>s, 100 kDa) to remove remaining iron<br />

salts, NH4OH <strong>and</strong> sodium citrate. If required, fresh additions <strong>of</strong> the same<br />

volume <strong>of</strong> Fe salt solutions were carried out under the same conditions<br />

as previously specified to ensure the full magnetic coverage <strong>of</strong> fibrils <strong>and</strong><br />

to create a nanowire.<br />

Characterisation <strong>of</strong> magnetic 1D assemblies <strong>and</strong> nanowires: TEM images<br />

were obtained with a transmission electron microscope (Phillips CM-12)<br />

operating at an accelerating voltage <strong>of</strong> 120 kV as previously described.<br />

HR-TEM images <strong>and</strong> SAED patterns were obtained with a transmission<br />

7372<br />

electron microscope (Carl-Zeiss Libra 200 FE-EFTEM) operating at<br />

200 kV. X-ray diffraction experiments were carried out with a rotating<br />

anode X-ray generator (Siemens D5005). Twin Gçbel mirrors were used<br />

to produce a well-collimated beam <strong>of</strong> CuKa radiation (l = 1.5418 A). Xray<br />

diffraction patterns were recorded with an imaging plate detector<br />

(AXS F.Nr. J2–394). FTIR spectra <strong>of</strong> HSA in aqueous solutions were determined<br />

with a FTIR spectrometer (model IFS-66v from Bruker) with a<br />

horizontal ZnS ATR accessory. The spectra were obtained at a resolution<br />

<strong>of</strong> 2 cm 1 <strong>and</strong> generally 200 scans were accumulated to provide a reasonable<br />

signal-to-noise ratio. Magnetic susceptibility measurements were carried<br />

out with a SQUID magnetometer (Quantum Design MPMS5, San<br />

Diego, CA). Iron concentrations <strong>of</strong> each sample for SQUID, relaxivity<br />

<strong>and</strong> MRI were determined by inductively coupled plasma atomic emission<br />

spectroscopy (ICP-AES, Varian).<br />

Magnetic resonance measurements <strong>and</strong> MRI imaging: Transverse <strong>and</strong><br />

longitudinal relaxation times were also measured at 9.4 T (400 MHz) <strong>and</strong><br />

378C with a Bruker Biospin USR94/20 instrument (Ettlingen, Germany).<br />

An inversion-recovery pulse sequence was used to measure the longitudinal<br />

relaxation times with 16 inversion recovery times logarithmically<br />

spaced. Both Carr–Purcell–Meiboom–Gill (CPMG) <strong>and</strong> spin-echo pulse<br />

sequences were used to measure transverse relaxation times. MRI phantoms<br />

were constructed in agarose solutions (Sigma–Aldrich, 4%, w/v),<br />

heated <strong>and</strong> stirred at 80 8C until complete solution <strong>of</strong> the solid agarose.<br />

Then, with the system still remaining fluid, six HPLC vials (Cromlab,<br />

Spain, 250 mL) were filled with the agarose solution (100 mL) <strong>and</strong> allowed<br />

to cool to room temperature. Protein solution (5, 10, 15, 25, 50 or<br />

100 mL) was then deposited over the surfaces <strong>of</strong> the agarose gels, <strong>and</strong> distilled<br />

water (49, 45, 40, 35, 25 <strong>and</strong> 0 mL, respectively) was added to the<br />

solutions to make the final volumes even at 150 mL per vial. After a waiting<br />

period <strong>of</strong> at least 30 min, most <strong>of</strong> the deposited fluid had been absorbed<br />

by the agarose gels, after which semi-gelificated agarose solution<br />

(50 mL, 10 %, w/v, heated to 80 8C for solution <strong>and</strong> cooled to about 40–<br />

508C before addition) was added to each phantom to seal them. Imaging<br />

<strong>of</strong> the phantoms was performed with the previously described Bruker<br />

Biospin system with 440 mT m 1 gradients. A 3D-Gradient Echo image<br />

was acquired (T2*-weighting) with the following parameters: field <strong>of</strong><br />

view: 28.8 ”19.2 ”2 mm, matrix size: 288 ” 192 ”40 points giving a spatial<br />

resolution <strong>of</strong> 100 ”100 ”200 mm, echo time TE = 8 ms, repetition time<br />

TR=100 ms <strong>and</strong> flip angle = 30.<br />

Acknowledgements<br />

P. Taboada et al.<br />

The authors thank the Ministerio de Ciencia e Innovación (MICINN) for<br />

research project MAT 2010–17336, the Xunta de Galicia for research<br />

project INCITE09206020PR <strong>and</strong> for European Regional Development<br />

Funds (research project 2010/50), <strong>and</strong> the Fundación Ramón Areces for<br />

additional financial support. We acknowledge the contributions <strong>of</strong> B. Argibay,<br />

P. Ramos-Cabrer <strong>and</strong> J. Castillo from the Clinical Neurosciences<br />

Research Laboratory <strong>of</strong> the Clinical Hospital <strong>of</strong> Santiago de Compostela<br />

for the MR images acquired with the 9.4 T MR system.<br />

[1] Y. N. Xia, P. D. Yang, Y. G. Sun, Y. Y. Wu, B. Mayers, B. Gates,<br />

Y. D. Yin, F. Kim, Y. Q. Yan, Adv. Mater. 2003, 15, 353 – 389.<br />

[2] a) C. Ross, Annu. Rev. Mater. Res. 2001, 31, 203 –235; b) J. M. Kinsella,<br />

A. Ivanisevic, J. Phys. Chem. C 2008, 112, 3191 –3193; c) C.<br />

Goubault, P. Jop, M. Fermigier, J. Baudry, E. Bertr<strong>and</strong>, J. Bibetter,<br />

Phys. Rev. Lett. 2003, 91, 260 802; d) J.-H. Park, G. von Maltzahn, L.<br />

Zhang„ M. P. Schwartz, E. Ruoslathi, S. N. Bathia, M. J. Sailor, Adv.<br />

Mater. 2008, 20, 1630 –1635.<br />

[3] S. Minko, A. Kiriy, G. Gorodyska, M. Stamm, J. Am. Chem. Soc.<br />

2002, 124, 10192.<br />

[4] a) N. Duxin, F. T. Liu, H. Vali, A. Eisenberg, J. Am. Chem. Soc.<br />

2005, 127, 10063 – 10069; b) Y. Piao, J. Kim, H. B. Na, D. Kim, J. S.<br />

Baek, M. K. Ko, J. H. Lee, M. Shokouhimehr, T. Hyeon, Nat. Mater.<br />

2008, 7, 242 –247.<br />

www.chemeurj.org 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 2011, 17, 7366 – 7373


One-Dimensional Magnetic Nanowires<br />

[5] a) V. SalgueiriÇo-Maceira, M. A. Correa-Duarte, M. BaÇobre-<br />

López, M. Grzelczak, M. Farle, L. M. Liz-Marzµn, J. Rivas, Adv.<br />

Funct. Mater. 2008, 18, 616 – 621; b) M. L. Górzny, A. S. Walton,<br />

S. D. Evans, Adv. Funct. Mater. 2010, 20, 1295 –1300; c) K. Nakata,<br />

Y. Hu, O. Uzun, O. Bakr, F. Stellacci, Adv. Mater. 2008, 20, 4294 –<br />

4299; d) S. H. Sun, Adv. Mater. 2006, 18, 393 –403; e) L. Fu, X. G.<br />

Liu, Y. Zhang, V. P. Dravid, C. A. Mirkin, Nano Lett. 2003, 3, 757 –<br />

760; f) S. Palacin, P. C. Hidber, J. P. Bourgoin, C. Miramond, C.<br />

Fermon, G. M. Whitesides, Chem. Mater. 1996, 8, 1316 – 1325.<br />

[6] a) H. Wang, A. J. Patil, S. Petrov, S. Mann, M. A. Winnik, I. Manners,<br />

Adv. Mater. 2009, 21, 1805 – 1808; b) J. Fresnais, J.-F. Berret, B.<br />

Frka-Petesic, O. S<strong>and</strong>re, R. Perzynski, Adv. Mater. 2008, 20, 3877 –<br />

3881; c) K. Keren, M. Krueger, R. Girald, G. Ben-Yoseph, U. Sivan,<br />

E. Braun, Science 2002, 297, 72 –75.<br />

[7] a) L. W. Yin, Y. B<strong>and</strong>o, Y. C. Zhu, D. Goldberg, M. S. Li, Adv.<br />

Mater. 2004, 16, 929 –933; b) J. Sharma, R. Chiabra, A. Cheng, J.<br />

Brownell, Y. Liu, H. Yan, Science 2009, 323, 112 –116; c) T.<br />

Kodama, A. Jain, K. E. Goodson, Nano Lett. 2009, 9, 2005 – 2009.<br />

[8] a) N. C. Bigall, M. Reitzig, W. Naumann, P. Simon, K.-H. Van PØe,<br />

A. Eychmüller, Angew. Chem. 2008, 120, 7994 –7997; Angew. Chem.<br />

Int. Ed. 2008, 47, 7876 –7879; b) S. Banerjee, M. G. C. Kahn, S. S.<br />

Wong, Chem. Eur. J. 2003, 9, 1898 –1908; c) W. Q. Han, A. Zettl,<br />

Nano Lett. 2003, 3, 681 – 683; d) M. B. Dickerson, K. H. S<strong>and</strong>hage,<br />

R. R. Naik, Chem. Rev. 2008, 108, 4935 –4978.<br />

[9] S. S. Behrens, J. Mater. Chem. 2008, 18, 3788 –3798.<br />

[10] a) D. Walsh, L. Arcelli, T. Ikoma, J. Tanaka, S. Mann, Nat. Mater.<br />

2003, 2, 386 –390; b) J.-H. Park, G. von Maltzahn, L. Zhang, A. M.<br />

Derfus, D. Simberg, T. J. Harris, E. Ruoslathi, S. N. Bathia, M. J.<br />

Sailor, Small 2009, 5, 694 – 700.<br />

[11] a) S. J. Byrne, S. A. Corr, Y. K. Gun ko, J. M. Kelly, D. F. Brougham,<br />

S. Ghosh, Chem. Commun. 2004, 2560 – 2561; b) J. M. Kinsella, A.<br />

Ivanisevic, Langmuir 2007, 23, 3886 – 3890; c) Q. Gu, D. T. Haynie,<br />

Mater. Lett. 2008, 62, 3047 – 3050.<br />

[12] a) M. Allen, J. W. M. Bulte, L. Liepold, G. Basu, H. A. Zywicke,<br />

J. A. Frank, M. Young, T. Douglas, Magn. Reson. Med. 2005, 54,<br />

807 – 812; b) A. A. Martinez-Morales, N. G. Portney, Y. Zhang, G.<br />

Destito, G. Budak, E. Ozbay, M. Manchester, C. S. Ozkan, M.<br />

Ozkan, Adv. Mater. 2008, 20, 4816 –4820.<br />

[13] I. A. Banerjee, L. Yu, M. Shima, T. Yoshino, H. Takeyama, T. Matsunaga,<br />

H. Matsui, Adv. Mater. 2005, 17, 1128 – 1131.<br />

[14] a) R. Tsukamoto, M. Muraoka, M. Seki, H. Tabata, I. Yamashita,<br />

Chem. Mater. 2007, 19, 2389 –2391; b) M. Knez, A. M. Bittner, F.<br />

Boes, C. wege, H. Jeske, E. Maiß, K. Kern, Nano Lett. 2003, 3,<br />

1079 – 1082.<br />

[15] a) C. M. Dobson, Nature 2003, 426, 884 – 890; b) I. Cherny, E. Gazit,<br />

Angew. Chem. 2008, 120, 4128 –4136; Angew. Chem. Int. Ed. 2008,<br />

47, 4062 –4069; c) T. R. Jahn, S. E. Radford, Arch. Biochem. Biophys.<br />

2008, 469, 100 –117.<br />

[16] a) P. Taboada, S. Barbosa, E. Castro, V. Mosquera, J. Phys. Chem. B<br />

2006, 110, 20733 –20736; b) J. Juµrez, P. Taboada, V. Mosquera, Biophys.<br />

J.2009, 96, 2353 –2370; c) E. Frare, P. P. De Laureto, J. Zurdo,<br />

C. M. Dobson, A. Fontana, J. Mol. Biol. 2004, 340, 1153 –1165;<br />

d) A. P. Pawar, K. F. Dubai, J. Zurdo, F. Chiti, M. Vendruscolo,<br />

C. M. Dobson, J. Mol. Biol. 2005, 350, 379 –392; e) V. H. Lieu, J. W.<br />

Wu, S. S. S. Wang, C. H. Wu, Biotechnol. Prog. 2007, 23, 698 –706.<br />

[17] T. Peters, Jr., All About Albumin: Biochemistry Genetics, <strong>and</strong> Medical<br />

Applications, Academic Press, New York, 1996.<br />

[18] a) W. J. Crookes-Goodson, J. M. Slocik, R. N. Naik, Chem. Soc. Rev.<br />

2008, 37, 2403 –2412; b) N. Ostrov, E. Gazit, Angew. Chem. 2010,<br />

FULL PAPER<br />

122, 3082 – 3085; Angew. Chem. Int. Ed. 2010, 49, 3018 –3021; c) M.<br />

Malisauskas, R. Meskys, L. A. Morozova-Roche, Biotechnol. Prog.<br />

2008, 24, 1166 –1170.<br />

[19] L. Yu, I. A. Banerjee, M. Shima, K. Rajan, H. Matsui, Adv. Mater.<br />

2004, 16, 709 –712.<br />

[20] S. R. Bull, M. O. Guller, R. A. Bras, T. J. Meade, S. I. Stupp, Nano<br />

Lett. 2005, 5, 1–4.<br />

[21] S. Ghosh, A. Mukherjee, P. J. Sadler, S. Verma, Angew. Chem. 2008,<br />

120, 2249 –2253; Angew. Chem. Int. Ed. 2008, 47, 2217 – 2221.<br />

[22] Y. Sahoo, A. Goodarzi, M. T. Swihart, T. Y. Ohulchanskyy, N. Kaur,<br />

E. P. Furlani, P. N. Prasad, J. Phys. Chem. B 2005, 109, 3879 – 3885.<br />

[23] P. Dutta, A. Manivannan, M. S. Seehra, N. Shah, G. P. Huffman,<br />

Phys. Rev. B 2004, 70, 174 428.<br />

[24] I. Martínez-Mera, M. E. Espinosa-Pesqueira, R. PØrez-Hernµndez, J.<br />

Arenas-Alatorre, Mater. Letters. 2007, 61, 4447 –4451.<br />

[25] J. Wan, W. Cai, J. Feng, X. Meng, E. Liu, J. Mater. Chem. 2007, 17,<br />

1188 – 1192.<br />

[26] a) H. Wang, Q.-W. Chen, L.-X. Sun, H.-p. Qi, X. Yang, S. Zhou, J.<br />

Xiong, Langmuir 2009, 25, 7135 – 7139; b) T. J. Daou, G. Pourroy, S.<br />

BØgin-Colin, J. M. Gren›che, C. Ulhaq-Bouillet, P.- LegarØ, P. Bernhardt,<br />

C. Leuvrey, G. Rogez, Chem. Mater. 2006, 18, 4399 –43404.<br />

[27] a) A. F. Gross, M. R. Diehl, K. C. Beverly, E. K. Richman, S. H. Tolbert,<br />

J. Phys. Chem. B 2003, 107, 5475 –5482; b) D. Baldomir, D. Serantes,<br />

M. Pereiro, J. Botana, J. E. Arias, S. H. Masunaga, J. Rivas, J.<br />

Nanosci. Nanotechnol. 2010, 10, 2717 –2721.<br />

[28] D. H. Han, H. L. Wang, J. Luo, J. Magn. Magn. Mater. 1994, 136,<br />

176 – 182.<br />

[29] G. F. Goya, T. S. Berquó, F. C. Fonseca, M. P. Morales, J. Appl. Phys.<br />

2003, 94, 3520 –3528.<br />

[30] J. Qin, S. Laurent, Y. S. Jo, A. Roch, M. Mikhaylova, Z. M. Bhujwalla,<br />

R. N. Muller, M. Muhammed, Adv. Mater. 2007, 19, 1874 –1878.<br />

[31] Y. Wang, Y. W. Ng, Y. Chen, B. Shuter, J. Yi, J. Ding, S.-c. Wang, S.-<br />

S. Feng, Adv. Funct. Mater. 2008, 18, 308 – 318.<br />

[32] L. Zhang, Y. Zhang, J. Magn. Magn. Mater. 2009, 321, L15 – L20.<br />

[33] N. Bloembergen, E. M. Purcell, R. V. Pound, Phys. Rev. 1948, 73,<br />

679 – 712.<br />

[34] J. F. Berret, N. Schonbeck, F. Gazeau, D. El Kharrat, O. S<strong>and</strong>re, A.<br />

Vacher, M. Airiau, J. Am. Chem. Soc. 2006, 128, 1755 –1761.<br />

[35] P. Caravan, C. T. Farrar, L. Frullano, R. Uppal, Contrast Media Mol.<br />

Imaging 2009, 4, 89–100.<br />

[36] a) S. A. Corr, S. J. Byrne, R. Tekoriute, C. J. Meled<strong>and</strong>ri, D. F. Bougham,<br />

M. Lynch, C. Kerskens, L. O Dwyer, Y. K. Gun ko, J. Am.<br />

Chem. Soc. 2008, 130, 4214 – 4215; b) M. Y. Shin, S. I. Kim, S. J. Lee,<br />

K. E. Lee, S.-S. Han, D.-P. Jang, Y.-B. Kim, Z.-H. Cho, I. So, G. M.<br />

Spinks, Langmuir 2008, 24, 12 107–12 111.<br />

[37] H. B. Na, I. C. Song, T. Hyeon, Adv. Mater. 2009, 21, 2133 –2148.<br />

[38] J. H. Lee, Y. M. Huh, Y. Jun, J. Seo, J. Jang, H. T. Song, S. Kim, E. J.<br />

Cho, H. G. Yoon, J. S. Suh, J. Cheon, Nat. Med. 2006, 13, 95–99.<br />

[39] J. M. Perez, L. Josephson, T. O Loughlin, D. Hogemann, R. Weissleder,<br />

Nat. Biotechnol. 2002, 20, 816 –820.<br />

[40] a) C. N. Pace, F. Vajdos, L. Fee, G. Grimsley, T. Gray, Protein Sci.<br />

1995, 4, 2411 –2423; b) T. Knubovets, J. J. Osterhout, P. J. Connolly,<br />

A. M. Klibanov, Proc. Natl. Acad. Sci. USA 1999, 96, 1262 – 1267.<br />

Received: December 20, 2010<br />

Revised: March 7, 2011<br />

Published online: May 6, 2011<br />

Chem. Eur. J. 2011, 17, 7366 – 7373 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org 7373


6.7.- Relevant aspects on the HSA fibril formation<br />

I. HSA has the ability to self-assemble into amyloid-like aggregates under different<br />

solution conditions, i.e., under both physiological <strong>and</strong> acidic pH at elevated<br />

temperature <strong>and</strong> different ionic strengths. Under these solution conditions, the HSA<br />

native state is destabilized generating partially folded states that can aggregate to<br />

form fibrils.<br />

II. At physiological pH, fibrillation is progressively faster <strong>and</strong> more efficient in the<br />

presence <strong>of</strong> up to 50 mM NaCl due to electrostatic shielding. ThT fluorescence, CD, FT-<br />

IR <strong>and</strong> Trp fluorescence spectra confirm the structural changes in both tertiary <strong>and</strong><br />

secondary structure along the HSA fibrillation process. In this way, large extents <strong>of</strong> β-<br />

sheet structure at large salt concentrations are also corroborated from the analysis <strong>of</strong><br />

far UV-CD spectra.<br />

III. Under acidic conditions, a progressive enhancement <strong>of</strong> HSA fibrillation is observed as<br />

electrolyte concentration in solution increases.<br />

IV. The fibrillation process <strong>of</strong> HSA does not show a lag-phase growth, except at acidic pH<br />

in the absence <strong>of</strong> NaCl. The HSA fibrillation is a downhill process which does not<br />

require a highly organized <strong>and</strong> unstable nucleus, with a progressive increase <strong>of</strong> the β-<br />

sheet (up to 26%) <strong>and</strong> an unordered conformation at the expense <strong>of</strong> the α-helical<br />

conformation.<br />

V. The fibrils obtained show a curly morphology <strong>and</strong> differ in length. Besides,<br />

suprafibrillar assemblies (spherulites <strong>and</strong> fibrillar gels) formed by the protein human<br />

serum albumin under different solution conditions.<br />

VI. Upon incubation at 65 0 C at both acidic (pH 2.5) <strong>and</strong> physiological (pH 7.4) pH in the<br />

presence <strong>of</strong> different amounts <strong>of</strong> added electrolyte, suprafibrillar assemblies,<br />

spherulites <strong>and</strong> fibrillar gels, are formed under different solution conditions. Fibrillar<br />

gels are formed through intermolecular nonspecific association <strong>of</strong> amyloid fibrils.<br />

Meanwhile, at a pH close to the isoelectric point <strong>of</strong> HSA (pH 5.5), particulate gels were<br />

observed as a consequence <strong>of</strong> a faster protein aggregation, which does not allow the<br />

necessary structural reorganization to enable the formation <strong>of</strong> more ordered<br />

structures.<br />

VII. Within spherulites, fibrils display a radial arrangement around a disorganized protein<br />

core with sizes <strong>of</strong> several micrometers, as revealed by POM <strong>and</strong> confocal microscopy.<br />

VIII. Protein structure destabilization <strong>and</strong> subsequent aggregation (into amorphous or well-<br />

defined structured aggregates) were conditioned by the experimental pH,<br />

215


temperatures, <strong>and</strong> solvent polarity.<br />

IX. Ethanol regulated the hydrogen bonding extension, the attractive hydrophobic<br />

interactions, <strong>and</strong> the protein accessible surface area which, in turn, give rise to<br />

intermediate structural states prone to form aggregate.<br />

X. The HSA aggregation rate at physiological condition is faster than under acidic one<br />

owing to the pre-existing structural differences <strong>of</strong> the protein molecules in solution<br />

<strong>and</strong> to changes in the nature <strong>and</strong> strength <strong>of</strong> intermolecular interaction upon<br />

incubation.<br />

XI. Ethanol promoted an α-helix to β-sheet conformational transition at intermediate<br />

alcohol concentrations, whereas at low <strong>and</strong> high ethanol contents α-helix prevailed as<br />

the predominant structure at room temperature <strong>and</strong> physiological pH.<br />

XII. Electrostatic <strong>and</strong> hydrophobic interactions are important to control stacking <strong>of</strong> β-<br />

sheets favouring the formation <strong>of</strong> amyloid-like fibrils at physiological pH <strong>and</strong> high<br />

temperature.<br />

6.8.- Relevant aspects on the gold <strong>and</strong> magnetic nanowires<br />

I. We report a method to obtain metallic Au nanowires by using Lys protein fibrils as<br />

bioscaffolds. The metallic coverage <strong>of</strong> nan<strong>of</strong>ibrils could be easily controlled by<br />

changing the numbers <strong>of</strong> sequential additions <strong>of</strong> gold growth solution into the fibril<br />

solution overall already preformed Au nanocrystals. In this way, a continuous coverage<br />

<strong>of</strong> the fibril surface is obtained as a result <strong>of</strong> the growth <strong>of</strong> pre-existing nanoparticles<br />

<strong>and</strong> generation <strong>of</strong> new ones on the fibrils surface <strong>and</strong> subsequent fusion <strong>of</strong> adjacent<br />

as-synthesized nanoparticles.<br />

II. UV-Vis spectra results showed that the hybrid fibrils can successfully catalyze the<br />

reduction reaction from p-nitrophenol to p-aminophenol by NaBH4. Moreover, the<br />

fibril metallic fibrils provide larger reaction rates values 1.5-4 than other hybrid<br />

materials reported, previously, as for example glucose-reduced Au nanoparticle<br />

networks, PAMAM-dendrimer-supported spherical nanoparticles amongst others.<br />

III. On the other h<strong>and</strong>, we report a method for obtaining magnetic Fe3O4 nanowires based<br />

on the use <strong>of</strong> HSA <strong>and</strong> Lys protein fibrils as bioscaffolds for generation <strong>of</strong> a complete<br />

magnetic coating layer on the biotemplate surface by controlled Fe nanoprecipitation<br />

in situ in the presence <strong>of</strong> NH4OH.<br />

IV. Thanks to their large saturation magnetization, the 1D magnetic nanowires can serve<br />

216


as efficient MRI contrast agents, with transverse relaxivities larger than those obtained<br />

previously for other 1D nanostructures derived from different methods. Their<br />

capability as contrast agents was verified by obtaining MRI images at different Fe<br />

concentrations in phantoms.<br />

217


Conclusions<br />

The work has two parts: The first one was focused on the characterization <strong>of</strong> three block<br />

copolymers (EO12SO10, EO10SO10EO10, <strong>and</strong> EO137SO18EO137), meanwhile the second part was<br />

dedicated to the study <strong>of</strong> the fibrillation <strong>of</strong> the protein human serum albumin that was cover<br />

with gold <strong>and</strong> magnetic nanoparticles.<br />

The general conclusions are:<br />

1. - Characterization <strong>of</strong> the block copolymers EO12SO10, EO10SO10EO10, y EO137SO18EO137, in<br />

solution at the air/water <strong>and</strong> chlor<strong>of</strong>orm/water interfaces.<br />

I. For E12S10 dilute solutions, the coexistence <strong>of</strong> spherical micelles with vesicular<br />

structures has been observed. In addition, the spontaneous formation <strong>of</strong> vesicles by<br />

poly(oxystyrene)-poly(oxyethylene) block copolymers is reported for the first time, as<br />

confirmed by light scattering, polarized light microscopy, <strong>and</strong> transmission <strong>and</strong> cryo-<br />

scanning electron microscopy data.<br />

II. In dilute solution, the self-assembly <strong>of</strong> copolymer E10S10E10 leads to the formation <strong>of</strong><br />

elongated micelles as supported by light scattering <strong>and</strong> transmission electron<br />

techniques.<br />

III. In the case <strong>of</strong> copolymer E137S18E137, typical spherical micelles are observed. Tube<br />

inversion <strong>and</strong> rheological measurements were used to define the sol- s<strong>of</strong>t- <strong>and</strong> gel-<br />

hard boundaries <strong>of</strong> this copolymer.<br />

IV. E12S10 <strong>and</strong> E10S10E10 copolymers did not form gels in the concentration range analysed.<br />

However, only certain concentrations <strong>of</strong> copolymer E10S10E10 were analysed by<br />

rheometry. From experimental data, an upturn in the low-frequency range <strong>of</strong> the<br />

stress moduli was observed, denoting the existence <strong>of</strong> an emerging slow process,<br />

which was assigned to the formation <strong>of</strong> an elastic network.<br />

V. Block copolymers E12S10, E10S10E10 <strong>and</strong> E137S18E137 showed spontaneous adsorption at<br />

the air-water interface, which slowed down when the hydrophobicity <strong>of</strong> the molecule<br />

was increased. In contrast, at the chlor<strong>of</strong>orm/water interface no measurable effect is<br />

observed for copolymer E12S10 due to the slow diffusion <strong>of</strong> its chains to the interface in<br />

219


combination with a low bulk concentration; this slow diffusion is associated with the<br />

fact that chlor<strong>of</strong>orm is a good solvent for both E <strong>and</strong> S blocks.<br />

VI. Copolymer E137S18E137 displays an adsorption isotherm with the four classical regions<br />

representing the pancake, mushroom, brush <strong>and</strong> condensed states; the presence <strong>of</strong> a<br />

pseudo-plateau is attributed to the pancake-to-brush transition as E chains submerge<br />

into the aqueous sub-phase. On the other h<strong>and</strong>, for E12S10 <strong>and</strong> E10S10E10 copolymers<br />

only two regions are observed in their adsorption isotherms as a consequence <strong>of</strong> their<br />

low molecular weights, short S <strong>and</strong> E block lengths, <strong>and</strong> much larger S/E ratio. This<br />

involves the disappearance <strong>of</strong> the pseudo-plateau region due to the decrease in the<br />

fractional interfacial area occupied by EO segments.<br />

VII. According to AFM images, circular micelles are observed on Langmuir-Blodgett films <strong>of</strong><br />

the copolymers obtained at two surface transfer pressures. A decrease in micelle size<br />

<strong>and</strong> an increase in monolayer thickness are observed with increases in transfer<br />

pressure. Aggregation numbers derived from AFM images increase with the increase<br />

<strong>of</strong> the S weight fraction at a certain deposition pressure, which can be a consequence<br />

<strong>of</strong> stronger attractive interactions between the S blocks to avoid contact with the<br />

solvent.<br />

2. - On the HSA fibrillation:<br />

I. HSA has the ability to self-assemble into amyloid-like aggregates under different<br />

solution conditions, i.e., under both physiological <strong>and</strong> acidic pH at elevated<br />

temperature <strong>and</strong> different ionic strengths <strong>and</strong> solvent compositions. Under these<br />

solution conditions, the HSA native state is destabilized generating partially folded<br />

states that can aggregate to form fibrils.<br />

II. At physiological pH, fibrillation is progressively faster <strong>and</strong> more efficient in the<br />

presence <strong>of</strong> up to 50 mM NaCl due to electrostatic shielding. ThT fluorescence, CD, FT-<br />

IR <strong>and</strong> Trp fluorescence spectra confirm the structural changes in both tertiary <strong>and</strong><br />

secondary structure along the HSA fibrillation process. In this way, large extents <strong>of</strong> β-<br />

sheet structure at large salt concentrations are also corroborated from the analysis <strong>of</strong><br />

far UV-CD spectra.<br />

III. Under acidic conditions, a progressive enhancement <strong>of</strong> HSA fibrillation is observed as<br />

electrolyte concentration in solution increases.<br />

IV. The fibrillation process <strong>of</strong> HSA does not show a lag-phase growth, except at acidic pH<br />

in the absence <strong>of</strong> NaCl. The HSA fibrillation is a downhill process which does not<br />

220


equire a highly organized <strong>and</strong> unstable nucleus, with a progressive increase <strong>of</strong> the β-<br />

sheet (up to 26%) <strong>and</strong> an unordered conformation at the expense <strong>of</strong> the α-helical<br />

conformation.<br />

V. The fibrils obtained show a curly morphology <strong>and</strong> differ in length. Besides,<br />

suprafibrillar assemblies (spherulites <strong>and</strong> fibrillar gels) formed by the protein human<br />

serum albumin under different solution conditions.<br />

VI. Upon incubation at 65 0 C at both acidic (pH 2.5) <strong>and</strong> physiological (pH 7.4) pH in the<br />

presence <strong>of</strong> different amounts <strong>of</strong> added electrolyte, suprafibrillar assemblies,<br />

spherulites <strong>and</strong> fibrillar gels, are formed under different solution conditions. Fibrillar<br />

gels are formed through intermolecular nonspecific association <strong>of</strong> amyloid fibrils.<br />

Meanwhile, at a pH close to the isoelectric point <strong>of</strong> HSA (pH 5.5), particulate gels were<br />

observed as a consequence <strong>of</strong> a faster protein aggregation, which does not allow the<br />

necessary structural reorganization to enable the formation <strong>of</strong> more ordered<br />

structures.<br />

VII. Within spherulites, fibrils display a radial arrangement around a disorganized protein<br />

core with sizes <strong>of</strong> several micrometers, as revealed by POM <strong>and</strong> confocal microscopy.<br />

VIII. Protein structure destabilization <strong>and</strong> subsequent aggregation (into amorphous or well-<br />

defined structured aggregates) were conditioned by the experimental pH,<br />

temperatures, <strong>and</strong> solvent polarity.<br />

IX. Ethanol regulated the hydrogen bonding extension, the attractive hydrophobic<br />

interactions, <strong>and</strong> the protein accessible surface area which, in turn, give rise to<br />

intermediate structural states prone to form aggregate.<br />

X. The HSA aggregation rate at physiological condition is faster than under acidic one<br />

owing to the pre-existing structural differences <strong>of</strong> the protein molecules in solution<br />

<strong>and</strong> to changes in the nature <strong>and</strong> strength <strong>of</strong> intermolecular interaction upon<br />

incubation.<br />

XI. Ethanol promoted an α-helix to β-sheet conformational transition at intermediate<br />

alcohol concentrations, whereas at low <strong>and</strong> high ethanol contents α-helix prevailed as<br />

the predominant structure at room temperature <strong>and</strong> physiological pH.<br />

XII. Electrostatic <strong>and</strong> hydrophobic interactions are important to control stacking <strong>of</strong> β-<br />

sheets favouring the formation <strong>of</strong> amyloid-like fibrils at physiological pH <strong>and</strong> high<br />

temperature.<br />

3. - On the metal <strong>and</strong> magnetic nanowires :<br />

I. We report a method to obtain metallic Au nanowires by using Lys protein fibrils as<br />

221


ioscaffolds. The metallic coverage <strong>of</strong> nan<strong>of</strong>ibrils could be easily controlled by<br />

changing the numbers <strong>of</strong> sequential additions <strong>of</strong> gold growth solution into the fibril<br />

solution overall already preformed Au nanocrystals. In this way, a continuous coverage<br />

<strong>of</strong> the fibril surface is obtained as a result <strong>of</strong> the growth <strong>of</strong> pre-existing nanoparticles<br />

<strong>and</strong> generation <strong>of</strong> new ones on the fibrils surface <strong>and</strong> subsequent fusion <strong>of</strong> adjacent<br />

as-synthesized nanoparticles.<br />

II. UV-Vis spectra results showed that the hybrid fibrils can successfully catalyze the<br />

reduction reaction from p-nitrophenol to p-aminophenol by NaBH4. Moreover, the<br />

fibril metallic fibrils provide larger reaction rates values 1.5-4 than other hybrid<br />

materials reported, previously, as for example glucose-reduced Au nanoparticle<br />

networks, PAMAM-dendrimer-supported spherical nanoparticles amongst others.<br />

III. On the other h<strong>and</strong>, we report a method for obtaining magnetic Fe3O4 nanowires based<br />

on the use <strong>of</strong> HSA <strong>and</strong> Lys protein fibrils as bioscaffolds for generation <strong>of</strong> a complete<br />

magnetic coating layer on the biotemplate surface by controlled Fe nanoprecipitation<br />

in situ in the presence <strong>of</strong> NH4OH.<br />

IV. Thanks to their large saturation magnetization, the 1D magnetic nanowires can serve<br />

as efficient MRI contrast agents, with transverse relaxivities larger than those obtained<br />

previously for other 1D nanostructures derived from different methods. Their<br />

capability as contrast agents was verified by obtaining MRI images at different Fe<br />

concentrations in phantoms.<br />

222


References<br />

1. Hiemenz, Paul C. Polymer Chemistry: The basic consepts. New York: Marcel Dekker, Inc.,<br />

1984. 0-8247-7082-X.<br />

2. Ebewele, R. O. Polymer science <strong>and</strong> technology. New York: CRC Press, 2000.<br />

3. Sun, S. F. Physical Chemistry <strong>of</strong> Macromolecules: Basic principles <strong>and</strong> issues. New York: John<br />

Wiley & Sons, Inc., 2004. 0-471-28138-7.<br />

4. Berg, J. M.; Tymoczko, J. L.; Stryer, L.; Clarke, N. D. Biquímica. Barcelona: Editorial Reverté,<br />

S. A., 2003.<br />

5. Vicent, María C.; Álvarez, Silvia; Zaragozá, José L. Ciencia y tecología de Polímeros.<br />

Valencia : Universdiad Politécnica de Valencia, 2006.<br />

6. Beyond molecules: <strong>Self</strong>-assembly <strong>of</strong> mesoscopic <strong>and</strong> macroscopic components. Whitesides,<br />

G. M.; Boncheva, M. 2002, PNAS, Vol. 99, pp. 4769-4774.<br />

7. Rao, C. N. R.; Müller, A.; Cheetham, A. K. Nanomaterials chemistry: recent developments<br />

<strong>and</strong> new directions. Weinheim : WILEY-VCH, 2007.<br />

8. <strong>Self</strong>-assembling materials for therapeutic delivery. Branco, M. C.; Schneider, J. P. 2009, Acta<br />

Biomaterialia, Vol. 5, pp. 817-831.<br />

9. Directing the self-assembly <strong>of</strong> block copolymers. Darling, S. B. 2007, Prog. Pol. Sci., Vol. 32,<br />

pp. 1152-1204.<br />

10. Solution self-assembly <strong>of</strong> tailor-made macromolecular building blocks prepared by<br />

controlled radical polymerization techniques. Lutz, Jean F. 2006, Polym Int, Vol. 55, pp. 979-<br />

993.<br />

11. <strong>Self</strong>-<strong>Assembly</strong> at All Scales. Whitesides, George M.; Grzybowski, Bartosz. 2002, Science,<br />

Vol. 295, pp. 2418-2421.<br />

12. Rotello, V. M.; Thayumanavan, S. Molecular recognition<strong>and</strong> polymers: Control <strong>of</strong> polymer<br />

structure <strong>and</strong> self-assembly. Hoboken, New Jersey: John Wiley & Sons, Inc., 2008.<br />

13. Bottom-up Nanoelectronics. Hadley, Peter. Amsterdam: 34 European Microwave<br />

Conference, 2004, pp. 141-145.<br />

14. Technologies for nan<strong>of</strong>luidic systems: top-down vs. bottom-up—a review. Mijatovic, D.;<br />

Eijkel, J. C. T.; van den Berg, A. 2005, Lab Chip 5, Vol. 5, pp. 492-500.<br />

15. Hamley, I. W. Block copolymers in solutions: fundamentals <strong>and</strong> application. John Wiley &<br />

Sons, Ltd, 2005.<br />

223


16. Hiemenz, P. C.; Rajagopalan, R. Principles <strong>of</strong> Colloid <strong>and</strong> Surface Chemistry. New York:<br />

Marcel Dekker, 1997.<br />

17. Hamley, I. W. Introduction to s<strong>of</strong>t matter: synthetic <strong>and</strong> biological self-assembling<br />

materials. Chichester: John Wiley & Sons, Ltd, 2007.<br />

18. Petty, M. C. Langmuir-blodgett films: an introduction. Cambridge: Cambridge University<br />

Press, 1996.<br />

19. Adamson, A. W.; Gast, A. P. Physical chemistry <strong>of</strong> surfaces. 6th ed. New York: John Wiley &<br />

Sons, Inc., 1997.<br />

20. Dynamic surface tension <strong>and</strong> dilational viscoelasticity <strong>of</strong> adsorption layers <strong>of</strong> a<br />

hydrophobically modified chitosan. Babak, Valery G.; Desbrièresb, Jacques; Tikhonova,<br />

Vladimir E. 2005, Colloids <strong>and</strong> Surfaces A: Physicochem. Eng. Aspects, Vol. 255, pp. 119-130.<br />

21. Adsorption kinetics <strong>of</strong> hydrophobic polysoaps at the methylene chloride–water interface.<br />

Babak, Valery; Boury, Frank. 2004, Colloids <strong>and</strong> Surfaces A: Physicochem. Eng. Aspects, Vol.<br />

243, pp. 33-42.<br />

22. Dilational viscoelasticity <strong>and</strong> relaxation properties <strong>of</strong> interfacial electrostatic complexes<br />

between oppositely charged hydrophobic <strong>and</strong> hydrophilic polyelectrolytes. Babak, V. G.;<br />

Barosb, F.; Desbrièresd, J. 2008, Colloids <strong>and</strong> Surfaces B: Biointerfaces, Vol. 65, pp. 43-49.<br />

23. Schärtl, Wolfgang. Light Scattering from Polymer Solutions <strong>and</strong> Nanoparticles Dispersions.<br />

Berlin: Springer, 2007.<br />

24. Berne, B. J.; Pecora, R. DYNAMIC LIGHT SCATTERING: with application to chemistry,<br />

biology, <strong>and</strong> physics. New York: John Wiley & Sons, INC., 2000.<br />

25. Xu, R. Particle characterization: light scattering methods. New York: Kluwer Academic<br />

Publishers, 2002.<br />

26. Teraoka, I. Polymer Solutions: An Introduction to Physical Properties. Iwao Teraoka.<br />

Brookling: John Wiley & Sons, Inc, 2002.<br />

27. Some applications <strong>of</strong> light scattering in materials science. Holoubek, J. 2007, Journal <strong>of</strong><br />

Quantitative Spectroscopy & Radiative Transfer, Vol. 106, pp. 104-121.<br />

28. Uchegbu, I. F.; Schätzlein, A. G. Polymer in drug delivery. Boca Raton: Taylor & Francis,<br />

2006.<br />

29. Young, R. J.; Lovell, P. A. Introduction to polymers. London: Chapman & Hall, 1991.<br />

30. Huglins, M. B. Ligth scattering from polymer solution. New York: Plenum Press, 1972.<br />

31. Kokhanovsky, A. A. Light Scattering Reviews 4: single light scattering <strong>and</strong> radiative<br />

transfer. Chichester: Praxis Publishing, 2009.<br />

32. Static <strong>and</strong> dynamic light scattering <strong>of</strong> biological macromolecules: what can we learn?<br />

Murphy, R. M. 1997, Current Opinion in Biotechnology, Vol. 8, pp. 25-30.<br />

224


33. Wang, Z. L. HANDBOOK OF MICROSCOPY FOR NANOTECHNOLOGY. Boston: KLUWER<br />

ACADEMIC PUBLISHERS, 2005.<br />

34. Electron Diffraction Using Transmission Electron Microscopy. Bendersky, L. A.; Gayle, F. W.<br />

2001, J. Res. Natl. Inst. St<strong>and</strong>. Technol., Vol. 106, pp. 997-1012.<br />

35. van Dyck, S. A.; van L<strong>and</strong>uyt, D. J.; van Tendeloo, G. Electron Microscopy: Principles <strong>and</strong><br />

Fundamentals. VCH, 1997.<br />

36. Yao, N.; Wang, Z. L. HANDBOOK OF MICROSCOPY FOR NANOTECHNOLOGY. BOSTON:<br />

KLUWER ACADEMIC PUBLISHERS, 2005.<br />

37. Goldstein, J. I.; Newbury, D. E.; Echlin, P.; Joy, D. C.; Lyman, C. E.; Lifshin, E.; Sawyer, L.;<br />

Michael, J. R. Scanning Electron Microscopy <strong>and</strong> X-Ray Microanalysis. New York: Kluwer<br />

Academic/Plenum Publishers, 2003.<br />

38. Bowen, W. R.; Hilal, N. Atomic force microscopy in process engineering: an introduction to<br />

AFM for improved processes <strong>and</strong> products. Elsevier, 2009. pp. 1-24. Vol. 1.<br />

39. Valeur, B. Molecular fluorescence principles <strong>and</strong> application. Weinheim: Wiley-VCH, 2002.<br />

40. van Holde, K. E.; Johnson, W. C.; Ho, P. S. Principles <strong>of</strong> Physical Biochemistry. Upper Saddle<br />

River: Pearson Prentice-Hall, 2006.<br />

41. The binding <strong>of</strong> thi<strong>of</strong>lavin-T to amyloid fibrils: localisation <strong>and</strong> implications. Krebs, M. R. H.;<br />

Bromley, E. H. C.; Donald, A. M. 2005, Journal <strong>of</strong> Structural Biology, Vol. 149, pp. 30-37.<br />

42. Wallace, B. A.; Janes, R. W. Modern techniques for circular dichroism <strong>and</strong> synchrotron<br />

radiation ciicular dichroism spectroscopy. Amsterdam: IOS Press, 2009.<br />

43. A Holistic Approach for Protein Secondary Structure Estimation from Infrared Spectra in<br />

H2O Solutions. Vedantham, G.; Sparks, H. G.; Sane, S. U.; Tzannis, S.; Przybycien, T. M. 2000,<br />

Analytical Biochemistry, Vol. 285, pp. 33-49.<br />

44. Schr<strong>and</strong>er, B. Infrered <strong>and</strong> raman spectroscopy: methods <strong>and</strong> applications. Weinheim:<br />

VCH, 1995.<br />

45. Protein secondary structures in water from second-derivative amide I infrared spectra.<br />

Dong, A.; Huang, P.; Caughey, W. S. 1990, Biochemistry, Vol. 29, pp. 3303-3308.<br />

46. Secondary structures <strong>of</strong> proteins adsorbed onto aluminum hydroxide: Infrared spectroscopic<br />

analysis <strong>of</strong> proteins from low solution concentrations. Dong, A.; Jones, L. S.; Kerwin, B. A.;<br />

Krishnan, S.; Carpenter, J. F. 2006, Analytical Biochemistry, Vol. 351, pp. 282-289.<br />

47. Conformational studies <strong>of</strong> alanine-rich peptide using CD <strong>and</strong> FTIR spectroscopy. Bagínska,<br />

K.; Makowska, J.; Wiczk, W.; Kasprzykowski, F.; Chmurzynski, L. 2008, J. Pept. Sci., Vol. 14, pp.<br />

283-289.<br />

48. Hunt, B. J.; James, M. I. polymer Characterisation. New York: Marcel Dekker, 1993.<br />

225


49. Goodwin, J. W.; Hughes, R. W. Rheology for chemists: an introduction. Cambridge: RSC<br />

Plublishing, 2008.<br />

50. Guinebretière, R. X-ray Diffraction by Polycrystalline Materials. London: ISTE Ltd, 2007.<br />

51. He, B. B. Two-dimensional X-Ray diffraction. New Jersey: John Wiley & Sons, 2009.<br />

52. Sun, S. F. Physical Chemistry <strong>of</strong> Macromolecules: Basic principles <strong>and</strong> issues. John Wiley &<br />

Sons, Inc., 2004. 0-471-28138-7.<br />

53. He, B. B. Two-dimensional X-Ray diffraction. New Jersey : John Wiley & Sons, 2009.<br />

54. Buchner, J.; Kiefhaber, T. Protein folding h<strong>and</strong>book. Weinheim : WILEY-VCH, 2005.<br />

55. Skoog, D. A.; Holler, F. J.; Crouch, S. R. Principles <strong>of</strong> instrumental analysis. Canada:<br />

Thomson Brooks/Cole, 2007.<br />

56. Murphy, D. P. Fundamentals <strong>of</strong> light microscopy <strong>and</strong> electronic imaging. Sumerset: WILEY-<br />

LISS, 2001.<br />

57. A Novel Bottom-Viewed Inductively Coupled Plasma-Atomic Emission Spectrometry. Chan,<br />

G. C. Y.; Chan, W. T. 2004, Spectrochimica Acta Part B, Vol. 59, pp. 41-58.<br />

58. Clarke, J.; Braginski, A. I. The SQUID h<strong>and</strong>book. Weinheim : WILEY-VCH, 2004. Vol. II<br />

Applications <strong>of</strong> SQUIDs <strong>and</strong> SQUID systems.<br />

59. Magnetic Nanoparticles for Drug Delivery. Arruebo, M.; Fernández-Pacheco, R.; Ibarra, M.<br />

R.; Santamaría, J. 2007, Nanotoday, Vol. 2, pp. 22-32.<br />

60. Mørup, S.; Hansen, M. F. Superparamagnetic Particles. H. Kronmüller <strong>and</strong> S. Parkin.<br />

H<strong>and</strong>book <strong>of</strong> Magnetism <strong>and</strong> Advanced Magnetic Materials. John Wiley & Sons, 2007, Vol. 4.<br />

61. Andrä, W.; Häfeli, U.; Hergt, R.; Misri, R. Application <strong>of</strong> Magnetic Particles in Medicine <strong>and</strong><br />

Biology. [book auth.] H. Kronmüller <strong>and</strong> S. Parkin. H<strong>and</strong>book <strong>of</strong> Magnetism <strong>and</strong> Advanced<br />

Magnetic Materials. John Wiley & Sons, 2007, Vol. 4.<br />

62. Nikiforov, V. N.; Filinova, E. Y. Biomedical Applications <strong>of</strong> Magnetic Nanoparticles. [book<br />

auth.] S. P. Gubin. Magnetic Nanoparticles. Weinheim: WILEY-VCH, 2009, Vol. 10.<br />

63. Goyen, M. Real whole-body MRI: requirements, indications, perspectives. New York: The<br />

McGraw-Hill Companies, Inc, 2008.<br />

64. Stroman, P. W. Essentials <strong>of</strong> functional MRI. Boca Raton, London, New York : CRC Press,<br />

2011.<br />

65. Lecomm<strong>and</strong>oux, SébastienLazzari, Massimo; Liu, Guojun. An Introduction to Block<br />

Copolymer Applications: State-<strong>of</strong>-the-Art <strong>and</strong> Future Developments. Massimo Lazzari, Guojun<br />

Liu <strong>and</strong> Sébastien Lecomm<strong>and</strong>oux. Block Copolymers in Nanoscience. Germany: WILEY-VCH,<br />

2006.<br />

66. Hamley, Ian W. The Physics <strong>of</strong> Block Copolymers. Oxford: Oxford University Press, 1998.<br />

226


67. Polymers Get Organized. Bucknall , David G.; Anderson, Harry L. 2003, Science, Vol. 302,<br />

pp. 1904-1905.<br />

68. Aggregation <strong>of</strong> water-soluble block copolymers in aqueous solutions: Recent trends.<br />

Nakashima, Kenichi; Bahadur, Pratap. 2006, Advances in Colloid <strong>and</strong> Interface Science , Vols.<br />

123-126, pp. 75–96.<br />

69. Polydispersity <strong>and</strong> block copolymer self-assembly. Lynda, Nathaniel A.; Meulerb, Adam J.;<br />

Hillmyer, Marc A. 2008, Progress in Polymer Science, Vol. 33, pp. 875-893.<br />

70. Ozin, Ge<strong>of</strong>frey A.; Arsenault, André C. <strong>Self</strong>-Assemblig Block Copolymers. Ge<strong>of</strong>frey A. Ozin<br />

<strong>and</strong> André C. Arsenault. Nanochemistry: A chemical approach to nanomaterials. Toronto: RSC<br />

Publishing, 2005.<br />

71. Structural Properties <strong>of</strong> <strong>Self</strong>-assembled <strong>Polymeric</strong> Aggregates in Aqueous Solutions.<br />

Mortensen, Kell. 2001, Polym. Adv. Technol. , Vol. 12, pp. 2-22.<br />

72. <strong>Self</strong>-<strong>Assembly</strong> <strong>of</strong> Block Copolymer Micelles in an Ionic Liquid. He, Y.; Li, Z.; Simone, P.;<br />

Lodge, T. P. 2006, J. Am. Chem. Soc., Vol. 128, pp. 2745-2750.<br />

73. Hubbard, A. T. Encyclopedia <strong>of</strong> Surface <strong>and</strong> Colloid Science. New York: Marcel Dekker, Inc.,<br />

2002.<br />

74. Interaction <strong>of</strong> polymerizing systems with rubber <strong>and</strong> its homologs. Part II. Interaction <strong>of</strong><br />

rubber in the polymerization <strong>of</strong> methyl methacrylate <strong>and</strong> <strong>of</strong> styrene. Merrett, F. M. 1954,<br />

Trans. Faraday Soc., Vol. 50, pp. 759-767.<br />

75. A Supramolecularly Assisted Transformation <strong>of</strong> Block-Copolymer Micelles into Nanotubes.<br />

Kim, Sung H.; Nederberg, Frederick; Jakobs, Robert; Tan, Jeremy P. K.; Fukushima, Kazuki;<br />

Nelson, Alshakim ; Meijer, E. W.; Yang, Yi Y.;. 2009, Angew. Chem. Int. Ed., Vol. 48, pp. 4508-<br />

4512.<br />

76. <strong>Self</strong>-Assembled Block Copolymer Aggregates: From Micelles to Vesicles <strong>and</strong> their <strong>Biological</strong><br />

Applications. Blanazs, A.; Armes, S. P.; Ryan, A. J. 2009, Macromol. Rapid Commun., Vol. 30,<br />

pp. 267-277.<br />

77. Structural study on the micelle formation <strong>of</strong> poly(ethylene oxide)-poly(propylene oxide)poly(ethylene<br />

oxide) triblock copolymer in aqueous solution. Mortensen, K.; Pedersen, J. S.<br />

1993, Macromolecules, Vol. 26, pp. 805-812.<br />

78. Structure <strong>of</strong> (Deuterated PEO)−(PPO)−(Deuterated PEO) Block Copolymer Micelles As<br />

Determined by Small Angle Neutron Scattering. Goldmints, I.; Yu, G. E.; Booth, C.; Smith, K. A.;<br />

Hatton, T. A. 1999, Langmuir, Vol. 15, pp. 1651-1656.<br />

79. Small-Angle Neutron Scattering Investigation <strong>of</strong> the Temperature-Dependent Aggregation<br />

Behavior <strong>of</strong> the Block Copolymer Pluronic L64 in Aqueous Solution. Yang, L.; Alex<strong>and</strong>ridis, P.;<br />

Steyler, D. C.; Kositzia, M. J.; Holwarth, J. F. 2000, Langmuir, Vol. 16, pp. 8555-8561.<br />

80. A Supramolecularly Assisted Transformation <strong>of</strong> Block-Copolymer Micelles into Nanotubes.<br />

Kim, Sung H.; Nederberg, Frederick; Jakobs, Robert; Tan, Jeremy P. K.; Fukushima, Kazuki;<br />

227


Nelson, Alshakim ; Meijer, E. W.; Yang, Yi Y. 2009, Angew. Chem. Int. Ed., Vol. 48, pp. 4508-<br />

4512.<br />

81. Influence <strong>of</strong> Ionic Surfactants on the Aggregation <strong>of</strong> Poly(Ethylene Oxide)-Poly(Propylene<br />

Oxide)-Poly(Ethylene Oxide) Block Copolymers Studied by Differential Scanning <strong>and</strong> Isothermal<br />

Titration Calorimetry . da Silva, R. C.; Ol<strong>of</strong>sson, G.; Schillén, K.; Loh, W. 2002, J. Phys. Chem. B,<br />

Vol. 106, pp. 1239-1246.<br />

82. Association <strong>and</strong> Surface Properties <strong>of</strong> Poly(ethylene oxide)−Poly(styrene oxide) Diblock<br />

Copolymers in Aqueous Solution. Mai, S. M.; Booth, C.; Kelaraskis, A.; Haverdraki, V.; Ryan, A.<br />

J. 2000, Langmuir, Vol. 16, pp. 1681-1688.<br />

83. Association Properties <strong>of</strong> a Diblock Copolymer <strong>of</strong> Ethylene Oxide <strong>and</strong> Styrene Oxide in<br />

Aqueous Solution Studied by Light Scattering <strong>and</strong> Rheometry. Kelaraski, A.; Havedraki, V.;<br />

Rekatas, C. J.; Mai, S. M.; Attwood, D.; Booth, C.; Ryan, A. J.; Hamley, I. W.; Martini, I. 2001,<br />

Macromol. Chem Phys., Vol. 202, pp. 1345-1354.<br />

84. Micellization <strong>and</strong> Gelation <strong>of</strong> Triblock Copolymers <strong>of</strong> Ethylene Oxide <strong>and</strong> Styrene Oxide in<br />

Aqueous Solution. Yang, Z.; Crothers, M.; Ricardo, N. M. P. S.; Chaibundit, C.; Taboada, P.;<br />

Mosquera, V.; Kelarakis, A.; Havredaki, V.; Mart, L.; Collect, H. J.; Attwood, D.; Heatley, F.;<br />

Booth, C. 2003, Langmuir, Vol. 19, pp. 943-950.<br />

85. Katime, I. Química Física Macromolecular. Universidad del País Vasco : Servicio Editorial del<br />

País Vasco Euskal Herriko , 1994.<br />

86. Bilayers <strong>and</strong> Interdigitation in Block Copolymer Vesicles. Battaglia, G.; Ryan, A. J. 2005, J.<br />

Am. Chem. Soc., Vol. 127, pp. 8757-8764.<br />

87. Shear-Induced Formation <strong>of</strong> Multilamellar Vesicles (“Onions”) in Block Copolymers. Zipfel,<br />

J.; Lindner, M.; Tsinou, M.; Alex<strong>and</strong>ridis, P.; Richtering, W. 1999, Langmuir, Vol. 15, pp. 2599-<br />

2602.<br />

88. The evolution <strong>of</strong> vesicles from bulk lamellar gels. Battaglia, G.; Ryan, A. J. 2005, Nature<br />

Materials, Vol. 4, pp. 869-876.<br />

89. Spontaneous Vesicle Formation in a Block Copolymer System. Bryskhe, K.; Jansson, J.;<br />

Topgaard, D.; Schillén, K.; Olsson, U. J. 2004, J. Phys. Chem. B, Vol. 108, pp. 9710-9719.<br />

90. Association behavior <strong>of</strong> mixed triblock copoly(oxyalkylene)s (type EBE <strong>and</strong> ESE) in aqueous<br />

solution. Ricardo, N. M. P. S.; Chaibundit, C.; Yang, Z.; Attwood, D.; Booth, C. 2006, Langmuir,<br />

Vol. 22, pp. 1301-1306.<br />

91. Mixtures <strong>of</strong> triblock copolymers E62P39E62 <strong>and</strong> E137S18E137 potential for drug delivery<br />

from in situ gelling micellar formulations. Pinho, M. E. N.; Costa, F. M. L. L.; Filho, F. B. S.;<br />

Ricardo, N. M. P. S.; Yeates, S. G.; Attwood, D.; Booth, C. 2007, Int. J. Pharm., Vol. 328, pp. 95-<br />

98.<br />

92. Creighton, T. E. Proteins: structure <strong>and</strong> molecular. New York: Freeman, 1993.<br />

93. Nelson, D. L.; Cox, M. M. Principles <strong>of</strong> biochemistry. New York: Freeman, 2008.<br />

228


94. Are there Pathways for Protein Folding? Levinthal, C. 1968, J. Chim. Phys. Phys. Chim. Biol.,<br />

Vol. 65, pp. 44-45.<br />

95. Protein Folding <strong>and</strong> Unfolding at Atomic Resolution. Fersht, A. R.; Daggett, V. 2002, Cell ,<br />

Vol. 108, pp. 573-582.<br />

96. Underst<strong>and</strong>ing protein folding via free-energy surfaces from theory <strong>and</strong> experiment.<br />

Dinner, A. R.; Sali, A.; Smith, L. J.; Dobson, C. M.; Karplus, M. 2000, Trends Biochem. Sci. 25<br />

(2000) 331–339., Vol. 25, pp. 331-339.<br />

97. Navigating the Folding Routes. Wolynes, P. G.; Onuchic, J. N; Thirumalai, D. 1995, Science,<br />

Vol. 267, pp. 1619-1620.<br />

98. Folding versus aggregation: Polypeptide conformations on competing pathways. Jahn, T.<br />

R.; Radford, S. E. 2008, Archives <strong>of</strong> Biochemistry <strong>and</strong> Biophysics, Vol. 469, pp. 100-117.<br />

99. Towards complete descriptions <strong>of</strong> the free–energy l<strong>and</strong>scapes <strong>of</strong> proteins. Vendruscolo, M.;<br />

Dobson, C. M. 2005, Philos. Transact. A Math. Phys. Eng. Sci., Vol. 363, pp. 433-450.<br />

100. Protein Aggregation <strong>and</strong> Amyloid Fibril Formation by an SH3 Domain Probed by Limited<br />

Proteolysis. Polverino de Laureto, P.; Taddei, N.; Frare, E.; Capanni, C.; Costantini, S.; Zurdo,J.;<br />

Chiti, F.; Dobson, C. M.; Fontana, A. 2003, J. Mol. Biol., Vol. 334, pp. 129-141.<br />

101. Folding funnels, binding funnels, <strong>and</strong> protein function. Tsai, C. J.; Kumar, S.; Nussinov, R.<br />

1999, Protein Sci., Vol. 8, pp. 1181-1190.<br />

102. Kinetics <strong>and</strong> thermodynamics <strong>of</strong> amyloid fibril assembly. Wetzel, R. 2006, Acc. Chem. Res.,<br />

Vol. 39, pp. 671-679.<br />

103. The prot<strong>of</strong>ilament structure <strong>of</strong> insulin amyloid fibrils. Jimenez, J. L.; Nettleton, E. J.;<br />

Bouchard, M.; Robinson, C. V.; Dobson, C. M.; Saibil, H. L. 14, 2002, PNAS, Vol. 99.<br />

104. Hierarchical assembly <strong>of</strong> β2-microglobulin amyloid in vitro revealed by atomic force<br />

microscopy. Kad, N. M.; Myers, S. L.; Smith, D. P.; Smith, D. A.; Radford, S. E.; Thomson, N. H.<br />

2003, J. Mol. Biol., Vol. 330, pp. 785–797.<br />

105. Einige Bemerkungen über den vegetabilischen Faserst<strong>of</strong>f und sein Verhältniss zum<br />

Stärkemehl. Schleiden, M. J. 1838, Ann. Physik., Vol. 119, pp. 391-397.<br />

106. On the structural definition <strong>of</strong> amyloid fibrils <strong>and</strong> other polypeptide aggregates. Fändrich,<br />

M. 2007, Cell. Mol. Life Sci., Vol. 64, pp. 2066-2078.<br />

107. Amyloids: Not only pathological agents but also ordered nanomaterials. Cherny, I.; Gazit,<br />

E. 2008, Angew. Chem. Int., Vol. 47, pp. 4062-4069.<br />

108. Prediction <strong>of</strong> the aggregation propensity <strong>of</strong> proteins from the primary sequence:<br />

Aggregation properties <strong>of</strong> proteomes. Castillo, V.; Graña-Montes, R.; Sabate, R.; Ventura, S.<br />

2011, Biotechnol. J., Vol. 6, pp. 674-685.<br />

229


109. Unraveling the Mysteries <strong>of</strong> Protein Folding <strong>and</strong> Misfolding. Ecroyd, H.; Carver, J. A. 2008,<br />

IUBMB Life, Vol. 60, pp. 769-774.<br />

110. Mechanisms <strong>of</strong> amyloid fibril formation–focus on domain-swapping. Zerovnik, E.; Stoka,<br />

V.; Mirtic, A.; Guncar, G.; Grdadolnik, J.; Staniforth, R. A.; Turk, D.; Turk, V. 2011, FEBS<br />

Journal, Vol. 278, pp. 2263-2282.<br />

111. Solid-State NMR Study <strong>of</strong> Amyloid Nanocrystals <strong>and</strong> Fibrils Formed by the Peptide<br />

GNNQQNY from Yeast Prion Protein Sup35p. van der Wel, P. C. A.; Lew<strong>and</strong>owski, J. R.; Griffin,<br />

R. G. 2007, J. Am. Chem. Soc., Vol. 129, p. 5117.<br />

112. Structural integrity <strong>of</strong> β-sheet assembly. Marshall, K. E.; Serpell, L. C. 2009, Biochem. Soc.<br />

Trans., Vol. 37, pp. 671-676.<br />

113. Amyloids Fibrils in Bionanobiotechnology. Waterhouse, S. H.; Gerrad, J. A. 2004, Aust. J.<br />

Chem., Vol. 57, pp. 519-523.<br />

114. Design <strong>of</strong> model systems for amyloid formation: lessons for prediction <strong>and</strong> inhibition.<br />

Pastor, M. T.; Esteras-Chopo, A.; de la Paz, M. L. 2005, Current Opinion in Structural Biology,<br />

Vol. 15, pp. 57-63.<br />

115. A designed system for assessing how sequence affects to conformational transitions in<br />

proteins. Ciani, B.; Hutchinson, E. G.; Sessions, R. B.; Woolfson, D. N. 2002, J Biol Chem, Vol.<br />

277, pp. 10150-10155.<br />

116. Mutational analysis <strong>of</strong> designed peptides that undergo structural transition from -helix<br />

to -sheet <strong>and</strong> amyloid fibril formation. Takahashi, Y.; Ueno, A.; Mihara, H. 2000, Structure<br />

Fold Des., Vol. 8, pp. 915-925.<br />

117. Amyloid architecture: complementary assembly <strong>of</strong> heterogeneous combinations <strong>of</strong> three<br />

or four peptides into amyloid fibrils. Takahashi, Y.; Ueno, A.; Mihara, H. 2002, Chem. Bio.<br />

Chem., Vol. 3, pp. 637-642.<br />

118. Exploring amyloid formation by a de novo design. Kammerer, R. A.; Kostrewa, D.; Zurdo,<br />

J.; Detken, A.; Garcia-Echeverria, C.; Green, J. D.; Muller, S. A.; Meier, B. H.; Winkler, F. K.;<br />

Dobson, C. M. 2004, Proc. Natl. Acad. Sci., Vol. 101, pp. 4435-4440.<br />

119. Hacking the code <strong>of</strong> amyloid formation. Pastor, M. T.; Esteras-Chopo, A.; Serrano, L.<br />

2007, Prion, Vol. 1, pp. 1-14.<br />

120. Why are “natively unfolded” proteins unstructured under phisiologic condition? Uversky,<br />

V. N.; Gillespie, J. R.; Fink, A. L. 2000, PROTEINS: Structure, Function, <strong>and</strong> Genetics, Vol. 41, pp.<br />

415-427.<br />

121. Conformational constraints for amyloid fibrillation: the importance <strong>of</strong> being unfolded.<br />

Uversky, V. N.; Fink, A. L. 2004, Biochim. Biophys. Acta, Vol. 1698, pp. 131-153.<br />

122. Partial denaturation <strong>of</strong> transthyretin is sufficient for amyloid fibril formation in vitro.<br />

Colon, W.; Kelly, J. W. 1992, Biochemistry, Vol. 31, pp. 8654-8660.<br />

230


123. Designing conditions for in vitro formation <strong>of</strong> amyloid prot<strong>of</strong>ilaments <strong>and</strong> fibrils. Chiti, F.;<br />

Webster, P.; Taddei, N.; Clark, A.; Stefani, M.; Ramponi, G.; Dobson, G. M. 1999, Proc. Natl.<br />

Acad. Sci., Vol. 96, pp. 3590-3594.<br />

124. Partially Unfolded States <strong>of</strong> β2-Microglobulin <strong>and</strong> Amyloid Formation in Vitro. McParl<strong>and</strong>,<br />

V. J.; Kad, N. M.; Kalverda, A. P; Brown, A.; Kirwin-Jones, P.; Hunter, M. G.; Sunde, M.;<br />

Radford, E. S. 2000, Biochemistry, Vol. 32, pp. 8735-8746.<br />

125. <strong>Assembly</strong> <strong>of</strong> Amyloid Prot<strong>of</strong>ibrils via Critical Oligomers-A Novel Pathway <strong>of</strong> Amyloid<br />

Formation. Modler, A. J.; Gast, K.; Lutsch, G.; Damaschun, G. 2003, J. Mol. Biol., Vol. 325, pp.<br />

135-148.<br />

126. Competing Pathways Determine Fibril Morphology in the <strong>Self</strong>-assembly <strong>of</strong> β2-<br />

Microglobulin into Amyloid. Gosal, W. S.; Morten, I. J.; Hewitt, E. W.; Smith, D. A.; Thomson,<br />

N. H.; Radford, S. E. 2005, J. Mol. Biol., Vol. 351, pp. 850-864.<br />

127. Direct Observation <strong>of</strong> Oligomeric Species formed in the Early Stages <strong>of</strong> Amyloid Fibril<br />

Formation using Electrospray Ionisation Mass Spectrometry. Smith, A. M.; Jahn, T. R.;<br />

Ashcr<strong>of</strong>t, A. E.; Radford, S. E. 2006, J. Mol. Biol., Vol. 364, pp. 9-19.<br />

128. Amyloid β-protein (Aβ) assembly: Aβ40 <strong>and</strong> Aβ42 oligomerize through distinct pathways.<br />

Bitan, G.; Kirkitadze, M. D.; Lomakin, A.; Voller, S. S.; Benedek, G. B.; Teplow, D. B. 2003,<br />

Proc. Natl. Acad. Sci., Vol. 100, pp. 330-335.<br />

129. Small Molecule Inhibitors <strong>of</strong> Aggregation Indicate That Amyloid β Oligomerization <strong>and</strong><br />

Fibrillization Pathways Are Independent <strong>and</strong> Distinct. Necula, M.; Kayed, R.; Milton, S.; Glabe,<br />

C. G. 2007, J. Biol. Chem., Vol. 282, pp. 10311-10324.<br />

130. Thermodynamic features <strong>of</strong> the thermal unfolding <strong>of</strong> human serum albumin. Picó, G. A.<br />

1997, International Journal <strong>of</strong> <strong>Biological</strong> Macromolecules, Vol. 20, pp. 63-73.<br />

131. Atomic structure <strong>and</strong> chemistry <strong>of</strong> human serum albumin. He, X. M.; Carter, D. C. 1992,<br />

Vol. 358, pp. 209-215.<br />

132. Two-dimensional/attenuated total reflection infrared correlation spectroscopy studies on<br />

secondary structural changes in human serum albumin in aqueous solutions: pH-dependent<br />

structural changes in the secondary structures <strong>and</strong> in the hydrogen bondings <strong>of</strong>. Murayama, K.;<br />

Wu, Y.; Czarnik-Matusewicz, B.; Ozaki, Y. 2001, J. Phys. Chem. B, Vol. 105, pp. 4763-4769.<br />

133. Spectroscopic studies on the effect <strong>of</strong> temperature on pH-induced folded states <strong>of</strong> human<br />

serum albumin. Shaw, A. K.; Pal, S. K. 2008, Journal <strong>of</strong> Photochemistry <strong>and</strong> Photobiology B:<br />

Biology, Vol. 90, pp. 69-77.<br />

134. Unfolding <strong>of</strong> acrylodan-labeled human serum albumin probed by steady-state <strong>and</strong> timeresolved<br />

fluorescence methods. Flora, K.; Brennan, J. D.; Baker, G. A.; Doody, M. A.; Bright, F.<br />

V. 1998, Biophysical Journal, Vol. 75, pp. 1084–1096.<br />

135. Thermal stability <strong>of</strong> bovine serum albumin DSC study. Michnik, A. 2003, Journal <strong>of</strong><br />

Thermal Analysis <strong>and</strong> Calorimetry, Vol. 71, pp. 509-519.<br />

231


136. Dilatational Rheology <strong>of</strong> BSA Conformers at the Air/Water Interface. Cascão Pereira, L. G.;<br />

Theódoly, O.; Blanch, H. W.; Radke, C. J. 2003, Langmuir, Vol. 19, pp. 2349-2356.<br />

137. The three recombinant domains <strong>of</strong> human serum albumin: structural characterization <strong>and</strong><br />

lig<strong>and</strong> binding properties. Dockal, M.; Carter, D. C.; Rüker, F. 41, 1999, The journal <strong>of</strong><br />

biological chemistry, Vol. 274, pp. 29303–29310.<br />

138. The behavior <strong>of</strong> polyamino acids reveals an inverse side chain effect in amyloid structure<br />

formation. Fändrich, M.; Dobson, C. M. 2002, The EMBO J., Vol. 21, pp. 5682–5690.<br />

139. Existence <strong>of</strong> different structural intermediates on the fibrillation pathway <strong>of</strong> human serum<br />

albumin. Juárez, J.; Taboada, P.; Mosquera, V. 2009, Biophysical Journal, Vol. 96, pp. 2353-<br />

2370.<br />

140. Influence <strong>of</strong> electrostatic interactions on the fibrillation process <strong>of</strong> human serum albumin.<br />

Juárez, J.; Taboada, P.; Mosquera, V. 2009, J. Phys. Chem. B, Vol. 113, pp. 10521-10529.<br />

141. Additional supra-self-assembly <strong>of</strong> human serum albumin under amyloid-like-forming<br />

solution conditions. Juárez, J.; Taboada, P.; Goy-López, S.; Cambón, Adriana; Madec, M. B.;<br />

Yeates, S. G.; Mosquera, V. 2009, J. Phys. Chem. B, Vol. 113, pp. 12391-12399.<br />

142. Electrochemical devices made from conducting nanowire networks self-assembled from<br />

amyloid fibrils <strong>and</strong> alkoxysulfonate PEDOT. Hamedi, M.; Herl<strong>and</strong>, A.; Karlsson, R. H.; Inganäs,<br />

O. 2008, Nano Lett., Vol. 8, pp. 1736-1740.<br />

143. Nanomaterials: amyloids reflect their brighter side. Mankar, S.; Anoop, A.; Sen, S.; Maji,<br />

S. K. 2011, Nano Reviews, Vol. 2, pp. 1-12.<br />

144. Nanomechanics <strong>of</strong> functional <strong>and</strong> pathological amiloid materials. Knowles, T. P. J.;<br />

Buehler, M. J. 2011, Nature nanotechnology, Vol. 6, pp. 469-479.<br />

145. Seeding-induced self-assembling protein nanowires dramatically increase the sensitivity <strong>of</strong><br />

immunoassays. Men, D.; Guo, Y. C.; Zhang, Z. P.; Wei, H. P.; Zohug, Y. F.; Cui, Z. Q.; Liang, X.<br />

S.; Li, K.; Leng, Y.; You, X. Y.; Zhang, X. E. 2008, Nano Lett., Vol. 9, pp. 2246-2250.<br />

146. Prot<strong>of</strong>ilaments, filaments, ribbons, <strong>and</strong> fibrils from peptidomimetic self-assembly:<br />

implications for amyloid fibril formation <strong>and</strong> materials science. Lashuel, H. A.; LaBrenz, S. R.;<br />

Woo, L.; Serpell, L. C.; Kelly, J. W. 2000, J. Am. Chem. Soc., Vol. 12, pp. 5262-5277.<br />

147. Ultrathin silver nanowires produced by amyloid biotemplating. Malisauskas, M.; Meskys,<br />

R.; Morozova-Roche, L. A. 2008, Biotechnol. Prog., Vol. 2004, pp. 1166-1170.<br />

148. LSPR-based nanobiosensors. Sepúlveda, B.; Angelomé, P. C.; Lechuga, L. M.; Liz-Marzán,<br />

L. M. 2009, Nano today, Vol. 4, pp. 244-251.<br />

149. Shape-controlled synthesis <strong>of</strong> silver nanoparticles for plasmonic <strong>and</strong> sensing applications.<br />

Cobley, C. M.; Skrabalak, S. E.; Xia, Y. 2009, Plasmonic, Vol. 4, pp. 171-191.<br />

150. A comparison study <strong>of</strong> the catalytic properties <strong>of</strong> Au-based nanocage, nanoboxes, <strong>and</strong><br />

nanoparticles. Zeng, J.; Zhang, Q.; Chen, J.; Xia, Y. 2010, Nano lett., Vol. 10, pp. 30-35.<br />

232


151. SERS-based diagnosis <strong>and</strong> biodetection. Alvarez-Puebla, R. A.; Liz-Marzán, L. M. 2010,<br />

Small, Vol. 6, pp. 604-610.<br />

152. A refined solution structure <strong>of</strong> hen lysozyme determined using residual dipolar coupling<br />

data. Schwalbe, H.; Grimshaw, S. B.; Spencer, A.; Buck, M.; Boyd, J.; Dobson, C. M.; Redfield,<br />

C.; Smith, L. J. 2001, Protein Science, Vol. 10, pp. 677-688.<br />

153. Amyloid fibril formation <strong>of</strong> hen lysozyme depends on the instability <strong>of</strong> the C-helix (88-99).<br />

Harada, A.; Azakami, H.; Kato, A. 2008, Biosci. Biotechnol. Biochem., Vol. 72, pp. 1523-1530.<br />

154. Comparative study <strong>of</strong> inactivation <strong>and</strong> conformational change <strong>of</strong> lysozyme induced by<br />

pulsed electric fields <strong>and</strong> heat. Zhao, W.; Yang, R. 2008, Eur. Food. Res. Technol., Vol. 228, pp.<br />

47-54.<br />

155. Thermally induced fibrillar aggregation <strong>of</strong> hen egg white lysozyme. Arnaudov, L. N.; de<br />

Vries, R. 2005, Biophysical Journal, Vol. 88, pp. 515-526.<br />

156. Catalytic properties <strong>of</strong> gold nanoparticles immobilized on the surfaces <strong>of</strong> Nanocarriers.<br />

Chen, X.; Zhao, D.; An, Y.; Shi, L.; Hou,W.; Chen, L. 2010, J. Nanopart. Res., Vol. 12, pp. 1877-<br />

1887.<br />

157. Preparation <strong>of</strong> gold-dendrimer nanocomposites by laser irradiation <strong>and</strong> their catalytic<br />

reduction <strong>of</strong> 4-nitrophenol, immobilization <strong>and</strong> Recovery <strong>of</strong> Au Nanoparticles from anion<br />

exchange resin. Hayakawa, K.; Yoshimura, T.; Esumi, K. 2003, Langmuir, Vol. 19, pp. 5517-<br />

5521.<br />

158. A nanoreactor framework <strong>of</strong> a Au@SiO2 yolk/shell structure for catalytic reduction <strong>of</strong> pnitrophenol.<br />

Lee, J.; Park, L. C.; Song, H. 2008, Adv. Mater. 2008, 20, 1523–1528., Vol. 20, pp.<br />

1523-1528.<br />

159. Responsive catalysis <strong>of</strong> thermoresponsive micelle-supported gold nanoparticles. Wang, Y.;<br />

Wei, G.; Zhang, W.; Jiang, X.; Zheng, P.; Shi, L.; Dong, A. 2007, J. Mol. Catal. A: Chem., Vol.<br />

266, pp. 233-238.<br />

160. Magnetic nanoparticle assemblies on denatured DNA show unusual magnetic relaxivity<br />

<strong>and</strong> potential applications for MRI. Byrne, S. J.; Corr, S. A.; Gun'ko, Y. K.; Kelly, J. M.;<br />

Brougham, D. F.; Ghosh , S. 2004, Chem. Commun., pp. 2560-2561.<br />

161. Linear assemblies <strong>of</strong> magnetic nanoparticles as MRI contrast agents. Corr, S. A.; Byrne, S.<br />

J.; Tekoriute, R.; Meled<strong>and</strong>ri, C. J.; Bougham, D. F.; Lynch, M.; Kerskens, C.; O'Dwyer, L.;<br />

Gun'ko, Y. K. 2008, J. Am. Chem. Soc., Vol. 130, pp. 4214-4215.<br />

162. Wrap–bake–peel process for nanostructural transformation from bold beta-FeOOH<br />

nanorods to biocompatible iron oxide nanocapsules. Piao, Y.; Kim, J.; Na, H. B.; Kim, D.; Baek,<br />

J. S.; Ko, M. K.; Lee, J. H.; Shokouhimehr, M.; Hyeon, T. 2008, Nat. Matter., Vol. 7, pp. 242-<br />

247.<br />

233

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!